content
stringlengths 1
15.9M
|
---|
\section{Introduction}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f1.eps}
\caption{\label{fig:one}
Light curve of OGLE-2004-BLG-254.
The lower two panels show the overall shape of the light curve and residual
from the best-fit model. The upper two panels show the enlargement
of the peak region enclosed by a small box in the lower panel.
We note that a model light curve varies depending on an observed passband
due to the chromaticity caused by the finite-source effect. The presented
model curve is based on the passband of the first observatory in the list.
However, the residuals of the individual data sets are based on the model
curves of the corresponding passbands.
Colors of data points are chosen to match those of the labels of observatories
where data were taken.
The two dotted vertical lines in the upper panel represent the limb-crossing
start/end times.
The peak source magnification $A_{\rm P}$ is given in the upper panel.
}\end{figure}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f2.eps}
\caption{\label{fig:two}
Light curve of MOA-2007-BLG-176. Notations are same as in Fig.1
}\end{figure}
When an astronomical object (lens) is closely aligned with a background star
(source), the light from the source is deflected by the gravity of the lens,
resulting in brightening of the source star. The magnification of the source
flux is related to the projected lens-source separation by
\begin{equation}
A={u^2+2 \over u\sqrt{u^2+4}},
\label{eq1}
\end{equation}
where the separation $u$ is expressed in units of the angular Einstein radius
$\theta_{\rm E}$. The Einstein radius is related to the physical parameters
of the lens system by
\begin{equation}
\theta_{\rm E}=\left( \kappa M \pi_{\rm rel}\right)^{1/2};\qquad
\pi_{\rm rel}={\rm AU} \left( {1\over D_{\rm L}}-{1\over D_{\rm S}}\right),
\label{eq2}
\end{equation}
where $\kappa=4G/(c^2{\rm AU})=8.14$ mas ${M_{\odot}}^{-1}$, $M$ is the mass of
the lens, $\pi_{\rm rel}$ is the relative source-lens parallax, and
$D_{\rm L}$ and $D_{\rm S}$ are the distances to the lens and
source star, respectively. The relative motion between the source, lens, and
observer leads to light variation of the source star (lensing event).
The first microlensing events were detected by \citet{alcock93} and
\citet{udalski93} from the experiments conducted based on the proposal of
\citet{paczynski86}. With the development of observational strategy combined
with upgraded instrument, the detection rate of lensing events has been
dramatically increased from several dozen events per year during
the early phase of lensing experiments to more than a thousand events
per year in current experiments.
The magnification of source star flux increases as the lens approaches closer
to the source star. For a small fraction of events, the lens-source separation
is even smaller than the source radius and the lens passes over the surface of
the source star. These events are of scientific importance
due to various reasons.
First, a high-magnification event with a lens passing over a source star provides
a rare chance to measure the brightness profile of a remote star. For such an
event, in which the lens-source separation is comparable to the source size near
the peak of the event, different parts of the source star are magnified
by different amounts. The resulting lensing light curve
deviates from the standard form of a point-source event \citep{witt94, gould94,
nemiroff94, alcock97} and the analysis of the deviation enables to measure
the limb-darkening profile of the lensed star \citep{witt95, loeb95, valls98,
bryce02, heyrovsky03}. With the same principle, it is also possible to study
irregular surface structures such as spots \citep{heyrovsky00, han00, hendry02,
rattenbury02}.
\begin{deluxetable*}{ll}
\tablecaption{Events With Lenses Passing Over Source Stars\label{table:one}}
\tablewidth{0pt}
\tablehead{
\multicolumn{1}{c}{event} &
\multicolumn{1}{c}{reference}
}
\startdata
OGLE-2004-BLG-254 & \citet{cassan06} / this work \\
OGLE-2004-BLG-482 & \citet{zub11} \\
MOA-2006-BLG-130/OGLE-2006-BLG-437 & \citet{baudry11} / under analysis \\
OGLE-2007-BLG-050/MOA-2007-BLG-103 & \citet{batista09} \\
MOA-2007-BLG-176 & this work \\
OGLE-2007-BLG-224/MOA-2007-BLG-163 & \citet{gould09} \\
MOA-2007-BLG-233/OGLE-2007-BLG-302 & this work \\
OGLE-2008-BLG-279/MOA-2008-BLG-225 & \citet{yee09} \\
OGLE-2008-BLG-290/MOA-2008-BLG-241 & \citet{fouque10} \\
MOA-2009-BLG-174 & this work \\
MOA-2009-BLG-411 & \citet{fouque11} / under analysis \\
MOA-2010-BLG-311 & \citet{hung11} / under analysis \\
MOA-2010-BLG-436 & this work \\
MOA-2010-BLG-523 & \citet{gould11} / under analysis \\
MOA-2011-BLG-093 & this work \\
MOA-2011-BLG-274 & this work \\
OGLE-2011-BLG-0990/MOA-2011-BLG-300 & this work \\
OGLE-2011-BLG-1101/MOA-2011-BLG-325 & this work
\enddata
\end{deluxetable*}
Second, it is possible to measure the Einstein radius of the lens and the
relative lens-source proper motion. The light curve at the moment of the
entrance (exit) of the lens into (from) the source surface exhibits inflection
of the curvature. The duration of the passage over the source as
measured by the interval between the entrance and exit of the lens over the
surface of the source star is
\begin{equation}
\Delta t_{\rm T} = 2\sqrt{{\rho_\star}^2-{u_0}^2}\ t_{\rm E},
\label{eq3}
\end{equation}
where $\rho_{\star}$ is the source radius in units of $\theta_{\rm E}$
(normalized source radius), $u_0$ is the lens-source separation normalized by
$\theta_{\rm E}$ at the moment of the closest approach (impact parameter),
and $t_{\rm E}$ is the time scale for the lens to transit $\theta_{\rm E}$
(Einstein time scale). The impact parameter and the Einstein time scale are
measured from the overall shape of the light curve and the duration of the
event. With the known $u_0$ and $t_{\rm E}$ combined with the measured
duration of passage over the source, the normalized source radius is measured
from the relation (\ref{eq3}). With the additional information about the angular
source size, $\theta_{\star}$, then the Einstein radius and the lens-source
proper motion are measured as $\theta_{\rm E}=\theta_\star/\rho_\star$ and
$\mu= \theta_{\rm E}/t_{\rm E}$, respectively. For general lensing events,
the Einstein time scale is the only measurable quantity related to the physical
parameters of the lens. However, the time scale results from the combination of
3 physical parameters of the mass of the lens, $M$, the distance to the lens,
$D_{\rm L}$, and the lens-source transverse speed, $v$, and thus the information
about the lens is highly degenerate. The Einstein radius, on the other hand,
does not depend on $v$ and thus the physical parameters of the lens can be
better constrained. For a fraction of events with long time scales, it is
possible to additionally measure the lens parallax, $\pi_{\rm E}=\pi_{\rm rel}
/\theta_{\rm E}$, from the deviation of the light curve induced by the orbital
motion of the Earth around the Sun. With the Einstein radius and the lens
parallax measured, the physical parameters of the lens are uniquely determined
\citep{gould97}.
Third, high-magnification events are sensitive to planetary companions of
lenses. This is because a planet induces a small caustic near the primary lens
and a high-magnification event resulting from the source trajectory passing
close to the primary has a high chance to produce signals indicating the
existence of the planet \citep{griest98}. For an event with a lens passing over
a source star, the planetary signal is weakened by the finite-source effect
\citep{bennett96}. Nevertheless, two of the microlensing planets were discovered
through this channel: MOA-2007-BLG-400 \citep{dong09} and
MOA-2008-BLG-310 \citep{janczak10}.
Fourth, high-magnification events provide a chance to spectroscopically study
remote Galactic bulge stars. Most stars in the Galactic bulge are too faint
for spectroscopic observations even with large telescopes. However, enhanced
brightness of lensed stars of high-magnification events allows spectroscopic
observation possible, enabling population study of Galactic bulge stars
\citep{johnson08, bensby09, bensby11, cohen09, epstein10}.
In this work, we present integrated results of analysis for 14
high-magnification events with lenses passing over source stars that have
been detected since 2004. Among them, 8 events were newly analyzed and one
event was reanalyzed with additional data.
\begin{deluxetable*}{llllll}
\tabletypesize{\small}
\tablecaption{Observatories\label{table:two}}
\tablewidth{0pt}
\tablehead{
\multicolumn{1}{c}{event} &
\multicolumn{1}{c}{MOA} &
\multicolumn{1}{c}{OGLE} &
\multicolumn{1}{c}{$\mu$FUN} &
\multicolumn{1}{c}{PLANET} &
\multicolumn{1}{c}{RoboNet} \\
\multicolumn{1}{c}{(RA,DEC)$_{J2000}$} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{/MiNDSTEp}
}
\startdata
OGLE-2004-BLG-254 & & $\rm{LCO_I}$ & $\rm{CTIO_{I,V}}$ (39/5) & $\rm{Boyden_I}$ (74) & \\
($17^{\rm h}56^{\rm m}36^{\rm s}\hskip-2pt.20$,$-32^{\circ}33^{\prime}01^{\prime\prime}\hskip-2pt.80$) & & (377) & $\rm{FCO_N}$ (129) & $\rm{Canopus_I}$ (59) & \\
& & & & $\rm{SAAO_I}$ (112) & \\
\hline
MOA-2007-BLG-176 & $\rm{Mt. John_R}$ & & $\rm{Auckland_R}$ (68) & $\rm{Canopus_I}$ (26) & $\rm{Danish_I}$ (2) \\
($18^{\rm h}05^{\rm m}00^{\rm s}\hskip-2pt.41$,$-25^{\circ}47^{\prime}03^{\prime\prime}\hskip-2pt.69$) & (1388) & & $\rm{CTIO_{I,V}}$ (41/4) & $\rm{Steward_I}$ (4) & \\
& & & $\rm{FCO_N}$ (33) & & \\
& & & $\rm{Lemmon_I}$ (66) & & \\
& & & $\rm{VLO_N}$ (129) & & \\
\hline
MOA-2007-BLG-233 & $\rm{Mt. John_R}$ & $\rm{LCO_I}$ & $\rm{CTIO_{I,V}}$ (80/5) & $\rm{Canopus_{I,V}}$ (60/5) & $\rm{Danish_I}$ (125) \\
/OGLE-2007-BLG-302 & (645) & (628) & $\rm{FCO_N}$ (23) & $\rm{Perth_I}$ (23) & \\
($17^{\rm h}54^{\rm m}14^{\rm s}\hskip-2pt.86$,$-31^{\circ}11^{\prime}02^{\prime\prime}\hskip-2pt.65$) & & & $\rm{Lemmon_I}$ (19) & $\rm{SAAO_I}$ (80) & \\
& & & $\rm{SSO_N}$ (80) & & \\
\hline
MOA-2009-BLG-174 & $\rm{Mt. John_R}$ & & $\rm{Bronberg_N}$ (147) & $\rm{Canopus_I}$ (40) & $\rm{LT_R}$ (7) \\
($18^{\rm h}02^{\rm m}07^{\rm s}\hskip-2pt.60$,$-31^{\circ}25^{\prime}24^{\prime\prime}\hskip-2pt.20$) & (2189) & & $\rm{CAO_N}$ (111) & & \\
& & & $\rm{Craigie_N}$ (130) & & \\
& & & $\rm{CTIO_{I,V}}$ (286/7) & & \\
& & & $\rm{Kumeu_N}$ (90) & & \\
& & & $\rm{Possum_N}$ (60) & & \\
\hline
MOA-2010-BLG-436 & $\rm{Mt. John_R}$ & & & $\rm{SAAO_{I,V}}$ (14/3) & $\rm{FTS_I}$ (3) \\
($18^{\rm h}03^{\rm m}21^{\rm s}\hskip-2pt.68$,$-27^{\circ}38^{\prime}10^{\prime\prime}\hskip-2pt.74$) & (2581) & & & & \\
\hline
MOA-2011-BLG-093 & $\rm{Mt. John_R}$ & $\rm{LCO_I}$ & $\rm{CTIO_{I,V}}$ (76/21) & $\rm{Canopus_I}$ (254) & $\rm{FTN_I}$ (3) \\
($17^{\rm h}46^{\rm m}17^{\rm s}\hskip-2pt.83$,$-34^{\circ}20^{\prime}24^{\prime\prime}\hskip-2pt.76$) & (2247) & (292) & $\rm{PEST_N}$ (124) & & $\rm{FTS_I}$ (19) \\
\hline
MOA-2011-BLG-274 & $\rm{Mt. John_R}$ & $\rm{LCO_I}$ & $\rm{Auckland_R}$ (53) & & \\
($17^{\rm h}54^{\rm m}42^{\rm s}\hskip-2pt.34$,$-28^{\circ}54^{\prime}59^{\prime\prime}\hskip-2pt.26$) & (3447) & (76) & $\rm{CTIO_I}$ (4) & & \\
& & & $\rm{FCO_N}$ (16) & & \\
& & & $\rm{Kumeu_R}$ (49) & & \\
& & & $\rm{PEST_N}$ (15) & & \\
\hline
OGLE-2011-BLG-0990 & $\rm{Mt. John_R}$ & $\rm{LCO_I}$ & $\rm{OPD_I}$ (275) & $\rm{Canopus_I}$ (10) & \\
/MOA-2011-BLG-300 & (1708) & (3434) & $\rm{Possum_R}$ (23) & $\rm{SAAO_{I,V}}$ (95/6) & \\
($17^{\rm h}51^{\rm m}30^{\rm s}\hskip-2pt.29$,$-30^{\circ}17^{\prime}47^{\prime\prime}\hskip-2pt.60$) & & & & & \\
\hline
OGLE-2011-BLG-1101 & $\rm{Mt. John_R}$ & $\rm{LCO_I}$ & $\rm{Auckland_R}$ (60) & $\rm{Canopus_I}$ (98) & $\rm{FTN_I}$ (65) \\
/MOA-2011-BLG-325 & (609) & (192) & $\rm{CTIO_{I,V}}$ (126/12) & & $\rm{FTS_I}$ (145) \\
($18^{\rm h}03^{\rm m}31^{\rm s}\hskip-2pt.62$,$-26^{\circ}20^{\prime}39^{\prime\prime}\hskip-2pt.50$) & & & $\rm{Possum_R}$ (24) & & $\rm{LT_I}$ (27) \\
& & & $\rm{SSO_N}$ (107) & & \\
& & & $\rm{VLO_N}$ (113) & &
\enddata
\tablecomments{
Mt. John: Mt. John Observatory, New Zealand; LCO: Las Campanas Observatory,
Chile; Auckland: Auckland Observatory, New Zealand; Bronberg: Bronberg
Observatory, South Africa; CAO: CAO San Pedro Observatory, Chile; Cragie:
Craigie Observatory, Australia; CTIO: Cerro Tololo Inter-American Observatory,
Chile; FCO: Farm Cove Observatory, New Zealand; Kumeu: Kumeu Observatory, New
Zealand; Lemmon: Mt Lemmon Observatory, Arizona, USA; OPD: Observatorio do
Pico dos Dias, Brazil; PEST: Perth Exoplanet Survey Telescope, Australia;
Possum: Possum Observatory, New Zealand; SSO: Southern Stars Observatory,
French Polynesia; VLO: Vintage Lane Observatory, New Zealand;
Boyden: Boyden Observatory, South Africa; Canopus: Canopus Hill Observatory,
Tasmania, Australia; Perth: Perth Observatory, Australia; SAAO: South African
Astronomical Observatory, South Africa; Steward: Steward Observatory, Arizona,
USA; FTN: Faulkes North, Hawaii; FTS: Faulkes South, Australia; LT: Liverpool
Telescope, La Palma, Spain; Danish: Danish Telescope, European Southern Observatory,
La Silla, Chile. The subscription after each observatory represents the filter
used for observation and the value in parenthesis is the number of data points.
The filter ``N'' denotes that no filter is used.
}
\end{deluxetable*}
\section{Event Selection}
The sample of events in our analysis is selected under the definition of {\it a single-lens
event where the lens-source separation at the time of the peak magnification
is less than the radius of the source star, i.e. $u_0 < \rho_*$ and thus
the lens passes over the surface of the source star.} To obtain a sample of events, we
begin with searching for high-magnification events that have been detected since
2004. Events with lenses passing over source stars can be usually distinguished
by the characteristic features of their light curves near the peak. These features
are the inflection of the curvature at the moment when the finite source
first touches and completely leaves the lens and the round shape of the light
curve during the passage of the lens over the source. To be more objective
than visual inspection, we conduct modeling of all high-magnification events
with peak magnifications $A_{\rm P}\ge10$ to judge the qualification of events.
From these searches, we find that there exist 18 such events. Among them,
analysis results of 12 events were not published before. We learn that
4 unpublished events MOA-2006-BLG-130/OGLE-2006-BLG-437 \citep{baudry11},
MOA-2009-BLG-411 \citep{fouque11}, MOA-2010-BLG-523 \citep{gould11},
and MOA-2010-BLG-311 \citep{hung11} are under analysis by other researchers
and thus exclude them in our analysis. We note that there exist
4 known source-crossing events detected before 2004, including
MACHO Alert 95-30 \citep{alcock97}, OGLE sc26\_2218 \citep{smith03},
OGLE-2003-BLG-238 \citep{jiang04}, and OGLE-2003-BLG-262 \citep{yoo04}. We also
note that MOA-2007-BLG-400 \citep{dong09} and MOA-2008-BLG-310 \citep{janczak10}
exhibit characteristic features of source-crossing single-lens events but
we exclude them in the sample because the lenses of the events turned out to
have planetary companions.
In this work, we conduct analyses of 9 events. Among them, 8 events are
newly analyzed in this work. These events include MOA-2007-BLG-176,
MOA-2007-BLG-233/OGLE-2007-BLG-302, MOA-2009-BLG-174, MOA-2010-BLG-436,
MOA-2011-BLG-093, MOA-2011-BLG-274, OGLE-2011-BLG-0990/MOA-2011-BLG-300,
and OGLE-2011-BLG-1101/MOA-BLG-2011-325. For OGLE-2004-BLG-254, which was
analyzed before by \citet{cassan06}, we conduct additional analysis by
adding more data sets taken from CTIO and FCO. \footnote{Besides the data
sets listed in Table \ref{table:two}, there exists an additional data set taken
by using the Danish telescope. However, we do not use these data because
it has been shown by \citet{heyrovsky08} that the large scatter of the
data results in poor measurement of lensing parameters including the
limb-darkening coefficient.} In Table \ref{table:one}, we summarize
the status of analysis for the total 18 events that have been detected since 2004.
\section{Observation}
For almost all events analyzed in this work, the source-crossing part of
the light curve was densely covered. This was possible due to the
coordinated work of survey and follow-up observations. Survey groups issued
alerts of events. For a fraction of the events with high-magnifications,
additional alerts were issued. In other cases, follow-up teams issued
high-magnification alerts independently. The peak time of a high-magnification
event was predicted by real-time modeling based on the rising part of
the light curve. Finally, the peak was densely covered by many telescopes
that were prepared for follow-up observations at the predicted time of the peak.
For MOA-2010-BLG-436, the rising part of the light curve was not covered by
survey observations due to the short time scale of the event and thus
no alert was issued. Nevertheless, the event was positioned in a high frequency
field of the MOA survey and thus the peak was covered densely enough to be
confirmed as an event with the lens passing over the source.
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f3.eps}
\caption{\label{fig:three}
Light curve of MOA-2007-BLG-233/OGLE-2007-BLG-302. Notations are same as in
Fig.1
}\end{figure}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f4.eps}
\caption{\label{fig:four}
Light curve of MOA-2009-BLG-174. Notations are same as in Fig.1
}\end{figure}
Table \ref{table:two} shows the observatories of the telescopes that were
used for observations of the individual events along with the observed passbands
(marked as subscripts after the observatory names) and the numbers of data
points (values in parentheses). Also marked are the coordinates (RA,DEC) of
the events. Survey observations were conducted by MOA
and OGLE groups using the 1.8 m telescope at Mt. John Observatory in New
Zealand and the 1.3m Warsaw University telescope at Las Campanas Observatory
in Chile, respectively. Follow-up observations were carried out by $\mu$FUN,
PLANET, RoboNet, and MiNDSTEp groups using 22 telescopes located in 8 different
countries. These telescopes include 1.3 m SMARTS CTIO, 0.4 m CAO in Chile,
0.4 m Auckland, 0.4 m FCO, 0.4 m Possum, 0.4 m Kumeu, 0.4 m VLO in New Zealand,
1.0 m Lemmon in Arizona, USA, 0.4 m Bronberg in South Africa, 0.6 m Pico dos Dias in Brazil,
0.25 m Craigie, 0.3 m PEST in Australia, 0.28m SSO in French Polynesia, 1.0 m SAAO,
1.5 m Boyden in South Africa, 1.0 m Canopus, 0.6 m Perth in Australia,
1.5 m Steward in Arizona, USA, 2.0 m FTN in Hawaii, USA, 2.0 m FTS in Australia,
2.0 m LT in La Palma, Spain, and 1.54 m Danish in La Silla, Chile.
Reduction of data was conducted by using photometry codes that were developed
by the individual groups. The MOA and OGLE data were reduced by photometry
codes developed by \citet{bond01} and \citet{udalski03}, respectively, which are
based on Difference Image Analysis method \citep{alard98}. The $\mu$FUN data were
processed using a DoPHOT pipeline \citep{schechter93}. For PLANET and MiNDSTEp data,
a pySIS pipeline \citep{albrow09} is used. For RoboNet data, a DanDIA pipeline
\citep{bramich08} is used.
The error bars estimated from different observatories are rescaled so that
$\chi^2$ per degree of freedom becomes unity for the data set of each observatory
where $\chi^2$ is computed based on the best-fit model. According to this simple
scheme, however, we find a systematic tendency for some data sets that error bars near the
peak of a light curve are overestimated. We find that this is caused by the inclusion
of redundant data at the baseline in error normalization. In this case, the data at the baseline
greatly outnumber accurate data points near the peak and thus error-bar normalization
is mostly dominated by the baseline data. To minimize this systematics, we restrict
the range of data for error normalization not to be too wide so that error estimation
is not dominated by data at the baseline, but not to be too narrow so that lensing
parameters can be measured accurately. For the final data set used for modeling,
we eliminate data points lying beyond 3$\sigma$ from the best-fit model.
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f5.eps}
\caption{\label{fig:five}
Light curve of MOA-2010-BLG-436. Notations are same as in Fig.1
}\end{figure}
In Figure \ref{fig:one}-\ref{fig:nine}, we present the light curves of the
individual events. In each figure, the lower two panels show the overall
shape of the light curve and residual and the upper two panels show the
enlargement of the peak region of the light curve and residual. For each figure,
we mark the moments of the lens'entrance and exit of the source by two dotted
vertical lines. Also marked is the peak source magnification. We note that
the same color of data points is used for each observatory throughout the light
curves and colors of data points are chosen to match those of labels of
observatories. We note that the magnitude scale corresponds to one of the observatories
in the list, while data from the other observatories have adjusted blends and are
vertically shifted to match the first light curve. The choice of reference is
based on data from survey observation, i.e. OGLE and MOA data. If both OGLE
and MOA data are available, the OGLE data is used for reference.
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f6.eps}
\caption{\label{fig:six}
Light curve of MOA-2011-BLG-093. Notations are same as in Fig.1
}\end{figure}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f7.eps}
\caption{\label{fig:seven}
Light curve of MOA-2011-BLG-274. Notations are same as in Fig.1
}\end{figure}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f8.eps}
\caption{\label{fig:eight}
Light curve of OGLE-2011-BLG-0990/MOA-2011-BLG-300. Notations are same as in Fig.1
}\end{figure}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f9.eps}
\caption{\label{fig:nine}
Light curve of OGLE-2011-BLG-1101/MOA-2011-BLG-325. Notations are same as in Fig.1
}\end{figure}
\section{Modeling}
Modeling the light curve of each event is conducted by searching for a set
of lensing parameters that best describes the observed light curve.
For all events, the light curves appear to have a standard form except
the peak region and thus we start with a simple single-lens modeling.
The light curve of a standard single-lensing event is characterized by
3 parameters, that are the time of the closest lens-source approach, $t_0$,
the lens-source separation at that moment, $u_0$, and the Einstein time scale,
$t_{\rm E}$. Based on the initial solution, we refine the solution by
considering additional second-order effects.
To precisely describe the peak region of the light curve of an event with
a lens passing over a source, additional parameters are needed to describe
the deviation caused by the finite-source effect. To the first order
approximation, the finite-source effect is described by the normalized
source radius, $\rho_\star$. For more refined description of the deviation,
additional parameters of the limb-darkening coefficients, $u_\lambda$,
are needed to account for the variation of the deviation caused by
the brightness profile of the source star surface. With the coefficients,
the limb-darkening profile is modeled by the standard linear law
\begin{equation}
I=I_0{\left[1-u_\lambda\left(1-cos\phi\right)\right]},
\label{eq4}
\end{equation}
where $I_0$ is the intensity of light at the center of the stellar
disk and $\phi$ is the angle between the normal to the stellar surface and
the line of sight toward the observer.
For an event with a time scale comparable to the orbital period of the Earth,
the position of the observer changes by the Earth's orbital motion during
the event and the resulting light curve deviates from a symmetric standard form.
This parallax effect is described by 2 parameters of $\pi_{{\rm E},N}$ and
$\pi_{{\rm E},E}$, that represent the two components of the lens parallax
vector $\mbox{\boldmath $\pi$}_{\rm E}$ projected on the sky in the north and east equatorial
coordinates, respectively. The direction of the parallax vector corresponds to the
lens-source relative motion in the frame of the Earth. The size of the
parallax vector corresponds to the ratio of the Earth's orbit, i.e. 1 AU,
to the Einstein radius projected on the observer's plane.
For a high-magnification event, the peak of the light curve can exhibit
additional deviations if the lens has a companion. For a planetary companion
located near the Einstein ring of the primary lens or a binary companion
with a separation from the primary substantially smaller or larger than
the Einstein radius, a small caustic is induced near the location of the
primary lens. Then, the source trajectory of a high-magnification event
passing close to the primary lens has a high chance to approach the caustic,
resulting in a perturbation near the peak of the light curve. Description
of the perturbation induced by a lens companion requires 3 additional
parameters of the mass ratio between the binary lens components, $q$, the
projected separation in units of the Einstein radius, $s$, and the angle
between the source trajectory and the binary axis, $\alpha$.
For each event, we search for a solution of the best-fit lensing parameters by
minimizing $\chi^2$ in the parameter space. For the $\chi^2$ minimization,
we use a Markov Chain Monte Carlo (MCMC) method. We compute finite
magnifications by using the ray-shooting technique \citep{schneider86,
kayser86, wambsganss97}. In this method, rays are uniformly shot from the
image plane, bent according to the lens equation, and land on the source
plane. Then, a finite magnification is computed by comparing the number
densities of rays on the image and source planes. Precise computation of
finite magnifications by using this numerical technique requires a large
number of rays and thus demands heavy computation. To minimize computation,
we limit finite-magnification computation by using the ray-shooting method
only when the lens is close to the source. Once a solution of the parameters
is found, we estimate the uncertainties of the individual parameters based on
the chain of solutions obtained from MCMC runs.
\begin{deluxetable*}{lllllll}
\tablecaption{Best-fit Parameters\label{table:three}}
\tablewidth{0pt}
\tablehead{
\multicolumn{1}{c}{event} &
\multicolumn{1}{c}{$\chi^{2}/{\rm dof}$} &
\multicolumn{1}{c}{$t_0$ (HJD-2450000)} &
\multicolumn{1}{c}{$u_0$} &
\multicolumn{1}{c}{$t_{\rm E}$ (days)} &
\multicolumn{1}{c}{$\rho_\star$} &
\multicolumn{1}{c}{$\pi_{\rm E}$}
}
\startdata
OGLE-2004-BLG-254 & 1326/593 & 3166.8194 & 0.0046 & 13.23 & 0.0400 & -- \\
& & $\pm$0.0002 & $\pm$0.0008 & $\pm$0.05 & $\pm$0.0002 & -- \\
OGLE-2004-BLG-482 & 756.9/693 & 3235.7816 & 0.000 & 9.61 & 0.1309 & -- \\
& & $\pm$0.0007 & $\pm$0.002 & $\pm$0.02 & $\pm$0.0005 & -- \\
OGLE-2007-BLG-050 & 1760.5/1745 & 4221.9726 & 0.002 & 68.09 & 0.0045 & 0.12 \\
/MOA-2007-BLG-103 & & $\pm$0.0001 & $\pm$0.000 & $\pm$0.66 & $\pm$0.0001 & $\pm$0.03 \\
OGLE-2007-BLG-224 & -- & -- & 0.00029 & 6.91 & 0.0009 & 1.97 \\
/MOA-2007-BLG-163 & & -- & -- & $\pm$0.13 & $\pm$0.0002 & $\pm$0.13 \\
OGLE-2008-BLG-279 & -- & 4617.34787 & 0.00066 & 106.0 & 0.00068 & 0.15 \\
/MOA-2008-BLG-225 & & $\pm$0.00008 & $\pm$0.00005 & $\pm$0.9 & $\pm$0.00006 & $\pm$0.02 \\
OGLE-2008-BLG-290 & 2317.7/2015 & 4632.56037 & 0.00276 & 16.36 & 0.0220 & -- \\
/MOA-2008-BLG-241 & & $\pm$0.00027 & $\pm$0.0002 & $\pm$0.08 & $\pm$0.0001 & -- \\
\hline
OGLE-2004-BLG-254 & 785.2/784 & 3166.823 & 0.0111 & 12.84 & 0.0418 & -- \\
& & $\pm$0.001 & $\pm$0.0004 & $\pm$0.09 & $\pm$0.0004 & -- \\
MOA-2007-BLG-176 & 1756.0/1747 & 4245.056 & 0.0363 & 8.13 & 0.0590 & -- \\
& & $\pm$0.001 & $\pm$0.0005 & $\pm$0.07 & $\pm$0.0006 & -- \\
MOA-2007-BLG-233 & 1779.4/1757 & 4289.269 & 0.0060 & 15.90 & 0.0364 & -- \\
/OGLE-2007-BLG-302 & & $\pm$0.001 & $\pm$0.0002 & $\pm$0.05 & $\pm$0.0001 & -- \\
MOA-2009-BLG-174 & 2816.5/3051 & 4963.816 & 0.0005 & 64.99 & 0.0020 & 0.06 \\
& & $\pm$0.001 & $\pm$0.0001 & $\pm$0.61 & $\pm$0.0001 & +0.07-0.02 \\
MOA-2010-BLG-436 & 2599.4/2593 & 5395.791 & 0.0002 & 12.78 & 0.0041 & -- \\
& & $\pm$0.001 & $\pm$0.0002 & $\pm$1.08 & $\pm$0.0003 & -- \\
MOA-2011-BLG-093 & 3038.0/3024 & 5678.555 & 0.0292 & 14.97 & 0.0538 & -- \\
& & $\pm$0.001 & $\pm$0.0002 & $\pm$0.05 & $\pm$0.0002 & -- \\
MOA-2011-BLG-274 & 3657.7/3649 & 5742.005 & 0.0029 & 2.65 & 0.0129 & -- \\
& & $\pm$0.001 & $\pm$0.0001 & $\pm$0.06 & $\pm$0.0003 & -- \\
OGLE-2011-BLG-0990 & 5551.6/5540 & 5758.691 & 0.0151 & 6.70 & 0.0199 & -- \\
/MOA-2011-BLG-300 & & $\pm$0.001 & $\pm$0.0004 & $\pm$0.07 & $\pm$0.0003 & -- \\
OGLE-2011-BLG-1101 & 1562.6/1562 & 5823.574 & 0.0485 & 29.06 & 0.0979 & -- \\
/MOA-2011-BLG-325 & & $\pm$0.002 & $\pm$0.0005 & $\pm$0.11 & $\pm$0.0006 & --
\enddata
\tablecomments{
The parameters of the first 6 events are adopted from previously analyses and
those of the other 9 events are determined in this work. For OGLE-2004-BLG-254,
the event was reanalyzed by adding more data sets. The references of the
previous analyses are presented in Table \ref{table:one}.
}
\end{deluxetable*}
\begin{deluxetable*}{lllll}
\tabletypesize{\small}
\tablecaption{Source Parameters\label{table:four}}
\tablewidth{0pt}
\tablehead{
\multicolumn{1}{c}{event} &
\multicolumn{1}{c}{source type} &
\multicolumn{3}{c}{limb-darkening coefficients} \\
\multicolumn{1}{c}{} &
\multicolumn{1}{c}{$(\log g, T_{\rm eff})$} &
\multicolumn{1}{c}{$u_{\it V}$} &
\multicolumn{1}{c}{$u_{\it R}$} &
\multicolumn{1}{c}{$u_{\it I}$}
}
\startdata
OGLE-2004-BLG-254 & KIII & -- & 0.70$\pm$0.05 & 0.55$\pm$0.05 \\
OGLE-2004-BLG-482 & MIII & -- & 0.88$\pm$0.02 & 0.71$\pm$0.01 \\
OGLE-2007-BLG-050 & subgiant & -- & -- & -- \\
/MOA-2007-BLG-103 & & & & \\
OGLE-2007-BLG-224 & FV & -- & -- & -- \\
/MOA-2007-BLG-163 & & & & \\
OGLE-2008-BLG-279 & GV & -- & -- & -- \\
/MOA-2008BLG-225 & & & & \\
OGLE-2008-BLG-290 & KIII & 0.77$\pm$0.01 & 0.62$\pm$0.07 & 0.55$\pm$0.01 \\
/MOA-2008-BLG-241 & & & & \\
\hline
OGLE-2004-BLG-254 & KIII & -- & -- & 0.70$\pm$0.07 (OGLE) \\
& (2.0, 4750 K) & & & 0.56$\pm$0.10 (CTIO) \\
& & & & 0.69$\pm$0.10 (Boyden) \\
& & & & 0.78$\pm$0.09 (Canopus) \\
& & & & 0.55$\pm$0.06 (SAAO) \\
& & & & {\bf 0.61 (Claret 2000)} \\
MOA-2007-BLG-176 & KIII & -- & 0.53$\pm$0.04 (MOA) & 0.50$\pm$0.05 (CTIO) \\
& (2.0, 4500 K) & & 0.51$\pm$0.05 (Auckland) & 0.44$\pm$0.06 (Lemmon) \\
& & & {\bf 0.73 (Claret 2000)} & {\bf 0.63 (Claret 2000)} \\
MOA-2007-BLG-233 & GIII & -- & 0.56$\pm$0.02 (MOA) & 0.53$\pm$0.04 (OGLE) \\
/OGLE-2007-BLG-302 & (2.5, 5000 K) & & {\bf 0.68 (Claret 2000)} & 0.56$\pm$0.02 (CTIO) \\
& & & & 0.49$\pm$0.02 (Canopus) \\
& & & & {\bf 0.59 (Claret 2000)} \\
MOA-2009-BLG-174 & FV & -- & -- & 0.33$\pm$0.02 (CTIO) \\
& (4.5, 6750 K) & & & {\bf 0.46 (Claret 2000)} \\
MOA-2010-BLG-436 & -- & -- & 0.52$\pm$0.10 (MOA) & -- \\
& & & & \\
MOA-2011-BLG-093 & GIII & 0.69$\pm$0.05 (CTIO) & 0.55$\pm$0.04 (MOA) & 0.51$\pm$0.10 (OGLE) \\
& (3.0, 5500 K) & {\bf 0.70 (Claret 2000)} & {\bf 0.63 (Claret 2000)} & 0.58$\pm$0.04 (CTIO) \\
& & & & 0.51$\pm$0.03 (Canopus) \\
& & & & {\bf 0.54 (Claret 2000)} \\
MOA-2011-BLG-274 & GV & -- & 0.48$\pm$0.02 (Kumeu) & -- \\
& (4.0, 6000 K) & & 0.51$\pm$0.01 (Auckland) & \\
& & & {\bf 0.59 (Claret 2000)} & \\
OGLE-2011-BLG-0990 & -- & -- & -- & 0.56$\pm$0.04 (OPD) \\
/MOA-2011-BLG-300 & & & & \\
OGLE-2011-BLG-1101 & KIII & 0.89$\pm$0.14 (CTIO) & 0.77$\pm$0.08 (MOA) & 0.74$\pm$0.07 (OGLE) \\
/MOA-2011-BLG-325 & (2.0, 4250 K) & {\bf 0.83 (Claret 2000)} & {\bf 0.76 (Claret 2000)} & 0.81$\pm$0.07 (CTIO) \\
& & & & 0.77$\pm$0.06 (Canopus) \\
& & & & 0.78$\pm$0.05 (FTS) \\
& & & & {\bf 0.65 (Claret 2000)}
\enddata
\tablecomments{
The parameters of the first 6 events are adopted from previously analyses and
those of the other 9 events are analyzed in this work. For OGLE-2004-BLG-254,
the event was reanalyzed by adding more data sets. The references of the
previous analyses are presented in Table \ref{table:one}. The limb-darkening
coefficients, $u_{\lambda}$, are presented for the individual data sets used
for $u_{\lambda}$ measurements and they are compared with theoretical values
predicted by \citet{claret00}. Also presented are the adopted values of $\log g$
and $T_{\rm eff}$. The unit of the stellar surface gravity is cm/s$^2$
}
\end{deluxetable*}
\section{Result}
In Table \ref{table:three}, we present the lensing parameters of the
best-fit solutions of the individual events determined from modeling.
To provide integrated results of events with lenses passing over source
stars, we also provide solutions of events that were previously analyzed.
For OGLE-2004-BLG-254, we provide both solutions of the previous
analysis and this work for comparison.
For all events analyzed in this work, we are able to measure the
limb-darkening coefficients of source stars. In Table \ref{table:four},
we present the measured limb-darkening coefficients. We measure the
coefficients corresponding to the individual data sets covering the peak
of each light curve instead of the individual passbands. This is because
the characteristics of filters used for different telescopes are different
from one another even though they are denoted by a single representative band and
thus joint fitting of data measured in different filter systems may result in
erroneous measurement of limb-darkening coefficients \citep{fouque10}.
To compare with theoretical values, we also provide values of coefficients
predicted by \citet{claret00} for the Bessell $V$, $R$, and $I$ filters.
Also provided are the source types and the adopted values of $\log g$ and
$T_{\rm eff}$ where the typical uncertainties of the surface gravity and
the effective temperature are $\Delta(\log g)=0.5$ and $\Delta T_{\rm eff}=250$ K,
respectively. We adopt a solar metallicity. We note that the measured
coefficients are generally in good agreement with theoretical values,
$u_{\rm th}$. From the table, it is found that for 23 out of the total
29 measurements the measured coefficients are within 20\% range of the
fractional difference as measured by $f_u=(u-u_{\rm th})/u_{\rm th}$.
The cases of large differences with $f_u>20\%$ include $u_I$(Canopus)
for OGLE-2004-BLG-254, $u_R$(MOA), $u_R$(Auckland), and $u_I$(Lemmon)
for MOA-2007-BLG-176, $u_I$(CTIO) for MOA-2009-BLG-174, and $u_I$(CTIO)
for OGLE-2011-BLG-1101/MOA-2011-BLG-325. From the inspection of the individual
data points on the light curves, we find the major reasons for the large
differences between the measured and theoretical values are due to poor
coverage [$u_I$(Canopus) for OGLE-2004-BLG-254, $u_R$(MOA, Auckland) and
$u_I$(Lemmon) for MOA-2007-BLG-176, $u_I$(CTIO) for MOA-2009-BLG-174] or
poor data quality [$u_I$(CTIO) for OGLE-2011-BLG-1101/MOA-2011-BLG-325].
Other possible reason for differences from the predicted values include
differences of individual filters from the standard Bessel filters,
as well as differences in the method to compute the theoretical values
\citep{heyrovsky07}.
The source type of each event is determined based on the location of the
source in the color-magnitude diagram (CMD) of stars in the same field.
CMDs are obtained from CTIO images taken in ${\it V}$ and ${\it I}$
bands. To locate the lensed star in the CMD, it is required to measure
the fraction of blended light in the observed light curve. This is done
by including a blending parameter in the process of light curve modeling.
For MOA-2011-BLG-274, a CMD taken from CTIO is available but images were
taken after the event and thus we could not determine the source color and
magnitude by the usual method. Instead we employ the method of \citet{gould10}.
In this method, we first measure the source instrumental magnitudes
by fitting the OGLE ($I_{\rm OGLE}$) and PEST (unfiltered, $N_{\rm PEST}$)
data to the light curve model. We then align each of these data sets to CTIO ($V/I$)
using comparison stars, which effectively transforms $N_{\rm PEST}/I_{\rm OGLE}$ to
$(V/I)_{\rm CTIO}$. In figure \ref{fig:ten}, we present the CMDs of stars
in the fields of the individual events and the locations of source stars.\footnote{We
note that high-resolution spectra are available for some events
with lenses passing over source stars. These events are OGLE-2004-BLG-254
\citep{cassan06}, OGLE-2004-BLG-482 \citep{zub11}, OGLE-2007-BLG-050/MOA-2007-BLG-103
\citep{johnson11}, MOA-2009-BLG-174, MOA-2010-BLG-311, MOA-2010-BLG-523
\citep{bensby11}, and MOA-2011-BLG-093 \citep{mcgregor11}. For those who
are more interested in the source stars of these events, see the related
references.} For MOA-2010-BLG-436 and OGLE-2011-BLG-0990/MOA-2011-BLG-300,
there exists SAAO data taken in {\it I} and {\it V} bands, but the number
and quality of {\it V}-band data are not numerous and good enough to specify
the source type.
In Table \ref{table:five}, we present the measured Einstein radii. The
Einstein radius of each event is determined from the angular source radius,
$\theta_{\star}$, and the normalized source radius, $\rho_{\star}$, as
$\theta_{\rm E}=\theta_{\star}/\rho_{\star}$. The normalized source radius
is measured from modeling. To measure the angular source radius, we use the
method of \citet{yoo04}, where the de-reddened ${\it V}-{\it I}$ color is measured
from the location of the source in the CMD, ${\it V}-{\it I}$ is converted into
${\it V}-{\it K}$ using the relation of \citet{bessel88}, and then the angular
source radius is inferred from the ${\it V}-{\it K}$ color and the surface
brightness relation given by \citet{kervella04}. In this process, we use
the centroid of bulge clump giants as a reference for the calibration of
the color and brightness of a source under the assumption that the source
and clump giants experience the same amount of extinction and reddening.
We note that no CMD is available for MOA-2010-BLG-436 and
OGLE-2011-BLG-0990/MOA-2011-BLG-300 and thus the Einstein radius
is not provided. Also provided in Table \ref{table:five} are
the relative lens-source proper motions as measured by $\mu=\theta_{\rm E}/t_{\rm E}$.
\begin{deluxetable*}{lllll}
\tablecaption{Physical Lens Parameters\label{table:five}}
\tablewidth{0pt}
\tablehead{
\multicolumn{1}{c}{event} &
\multicolumn{1}{c}{$\theta_{\rm E} \left({\rm mas}\right)$} &
\multicolumn{1}{c}{$\mu \left({\rm mas}~{\rm yr}^{-1}\right)$} &
\multicolumn{1}{c}{$M \left(M_{\sun}\right)$} &
\multicolumn{1}{c}{$D_{\rm L} \left({\rm kpc}\right)$}
}
\startdata
OGLE-2004-BLG-254 & 0.114 & 3.1 & -- & -- \\
OGLE-2004-BLG-482 & 0.4 & 16 & -- & -- \\
OGLE-2007-BLG-050/MOA-2007-BLG-103 & 0.48$\pm$0.01 & 2.63$\pm$0.08 & 0.50$\pm$0.14 & 5.5$\pm$0.4 \\
OGLE-2007-BLG-224/MOA-2007-BLG-163 & 0.91$\pm$0.04 & 48$\pm$2 & 0.056$\pm$0.004 & 0.53$\pm$0.04 \\
OGLE-2008-BLG-279/MOA-2008-BLG-225 & 0.81$\pm$0.07 & 2.7$\pm$0.2 & 0.64$\pm$0.10 & 4.0$\pm$0.6 \\
OGLE-2008-BLG-290/MOA-2008-BLG-241 & 0.30$\pm$0.02 & 6.7$\pm$0.4 & -- & -- \\
\hline
OGLE-2004-BLG-254 & 0.14$\pm$0.01 & 4.06$\pm$0.35 & -- & -- \\
MOA-2007-BLG-176 & 0.14$\pm$0.01 & 6.21$\pm$0.54 & -- & -- \\
MOA-2007-BLG-233/OGLE-2007-BLG-302 & 0.17$\pm$0.01 & 3.81$\pm$0.33 & -- & -- \\
MOA-2009-BLG-174 & 0.43$\pm$0.04 & 2.40$\pm$0.24 & 0.84$\pm$0.37 & 6.39$\pm$1.11 \\
MOA-2010-BLG-436 & -- & -- & -- & -- \\
MOA-2011-BLG-093 & 0.07$\pm$0.01 & 1.80$\pm$0.16 & -- & -- \\
MOA-2011-BLG-274 & 0.08$\pm$0.01 & 11.18$\pm$0.97 & -- & -- \\
OGLE-2011-BLG-0990/MOA-2011-BLG-300 & -- & -- & -- & -- \\
OGLE-2011-BLG-1101/MOA-2011-BLG-325 & 0.24$\pm$0.02 & 2.99$\pm$0.26 & -- & --
\enddata
\tablecomments{
The parameters of the first 6 events are adopted from previously analyses and
those of the other 9 events are analyzed in this work. For OGLE-2004-BLG-254,
the event was reanalyzed by adding more data sets. The references of the
previous analyses are presented in Table \ref{table:one}.
}
\end{deluxetable*}
\begin{figure*}[ht]
\epsscale{0.8}
\plotone{f10.eps}
\caption{\label{fig:ten}
Color-magnitude diagrams of neighboring stars in the fields of lensing events.
In each panel, the circle represents the location of the lensed star
and the triangle is the centroid of clump giants that is used as a reference
for color and brightness calibration. For MOA-2009-BLG-174,
the `X' mark denotes the location of the blend.
}\end{figure*}
\begin{figure}[ht]
\epsscale{1.1}
\plotone{f11.eps}
\caption{\label{fig:eleven}
Contours of $\chi^2$ from the best-fit solution in the space of the parallax
parameters of the event MOA-2009-BLG-174.
}\end{figure}
\begin{figure*}[ht]
\epsscale{1.1}
\plotone{f12.eps}
\caption{\label{fig:twelve}
Exclusion diagrams of planets as a function of the star-planet separation
(normalized in the Einstein radius) and the planet/star mass ratio.
}\end{figure*}
We note that the measured Einstein radii of some events are substantially
smaller than a typical value. These events include OGLE-2004-BLG-254
($\theta_{\rm E}\sim 0.14~{\rm mas}$), MOA-2007-BLG-176
($\sim 0.14~{\rm mas}$), MOA-2007-BLG-233/OGLE-2007-BLG-302
($\sim 0.17~{\rm mas}$), MOA-2011-BLG-093 ($\sim 0.07~{\rm mas}$), and
MOA-2011-BLG-274 ($\sim 0.08~{\rm mas}$).
The lens mass and distance are related to the Einstein radius by
\begin{equation}
M = 0.019~M_\odot
\biggl({D_{\rm S}\over 8~{\rm kpc}}\biggr)
\biggl({D_{\rm L}\over {D_{\rm S} - D_{\rm L}}}\biggr)
\biggl({\theta_{\rm E}\over 0.14~{\rm mas}}\biggr)^2
\label{eq5}
\end{equation}
Hence, the small $\theta_{\rm E}$ of these events implies that lenses
are either very close to the source or very low-mass objects. Most of
these events have proper motions that are typical of bulge lenses
(2--7 $\rm mas\,yr^{-1}$) and so may be quite close to the source
(see Table 5). But MOA-2011-BLG-274 has a substantially higher proper motion,
$\mu\sim11\ \rm mas\ yr^{-1}$. It is therefore a good candidate for
a sub-stellar object or even a free-floating planet \citep{sumi11}.
Because of its high proper motion, it should be possible to detect the lens
within a few years using high-resolution infrared imaging, provided
it is luminous. In this case a null result would confirm its substellar nature.
For MOA-2009-BLG-174, the lens parallax is measured with $\Delta\chi^2\sim
16.2$. The measured parallax parameters are
\begin{equation}
\pi_{{\rm E},\parallel}=-0.049\pm0.006~;\qquad \pi_{{\rm E},\perp}=0.038\pm0.065,
\end{equation}
where $\pi_{{\rm E},\parallel}$ and $\pi_{{\rm E},\perp}$ are the components
of the lens parallax vector that are parallel with and perpendicular to the
projected position of the Sun. These values correspond to the standard parallax
components of $(\pi_{{\rm E},N},\pi_{{\rm E},E})=(0.025\pm0.052,-0.057\pm0.028)$.
In Figure \ref{fig:eleven}, we present contours of $\chi^2$ in the space of
the parallax parameters. Combined with the measured Einstein radius,
the physical parameters of the lens are uniquely determined as
\begin{equation}
M={{\theta_{\rm E}}\over{\kappa\pi_{\rm E}}}=0.84\pm0.37\ M_{\sun},
\label{eq6}
\end{equation}
and
\begin{equation}
D_{\rm L}={{\rm AU}\over{\pi_{\rm E}\theta_{\rm E}+\pi_{\rm S}}}
=6.39\pm1.11\ {\rm kpc},
\label{eq7}
\end{equation}
respectively. We find that the measured lens mass is consistent with the
de-reddened color of blended light $\left({\it V}-{\it I}\right)_{0,b}\sim 1.4$,
which approximately corresponds to the color of an early K-type main-sequence
star with a mass equivalent to the estimated lens mass, suggesting that the blend
is very likely to be the lens. We mark the position of the blend in the
corresponding CMD in Figure \ref{fig:ten}.
A high-magnification event is an important target for planet search due to
its high efficiency to planetary perturbations. Unfortunately, we find no
statistically significant deviations from the single-lens fit for any of
the events analyzed in this work. However, it is still possible to place
limits on the range of the planetary separation and mass ratio. For this
purpose, we construct so-call ``exclusion diagrams'' which show the confidence
levels of excluding the existence of a planet as a function of the normalized
star-planet separation and the planet/star mass ratio. We construct diagrams
by adopting \citet{gaudi00} method. In this method, binary models are fitted
to observed data with the 3 binary parameters ($s$, $q$, $\alpha$) are held fixed.
Then, the confidence level of exclusion for planets with $s$ and $q$ is estimated
as the fraction of binary models not consistent with the best-fit single-lens
model among all tested models with various values of $\alpha$. For fitting
binary models, it is required to produce many light curves with finite magnifications.
We produce light curves by using the ``map-making method'' \citep{dong06},
where a magnification map for a given $s$ and $q$ is constructed and light curves
with various source trajectories are produced based on the map. In Figure
\ref{fig:twelve}, we present the obtained exclusion diagrams for all analyzed
events. Here we adopt a threshold of planet detection as $\Delta \chi^2_{\rm th}
= \chi^2_{\rm s}-\chi^2_{\rm p} = 200$, where $\chi^2_{\rm p}$ and $\chi^2_{\rm s}$
represent the $\chi^2$ values for the best-fit planetary and single-lens
models, respectively. For most events, the constraints on the excluded parameter
space is not strong mainly due to the severe finite-source effect. However,
the constraint is strong for MOA-2009-BLG-174 because of the small
source size ($\rho_{\star} \sim 0.002$) and dense coverage of the peak.
\section{Summary}
We provide integrated results of analysis for 14 high-magnification lensing
events with lenses passing over the surface of source stars that have been
detected since 2004. Among them, 8 events are newly analyzed in this work.
The newly analyzed events are
MOA-2007-BLG-176, MOA-2007-BLG-233/OGLE-2007-BLG-302, MOA-2009-BLG-174,
MOA-2010-BLG-436, MOA-2011-BLG-093, MOA-2011-BLG-274,
OGLE-2011-BLG-0990/MOA-2011-BLG-300, and OGLE-2011-BLG-1101/MOA-2011-BLG-325.
Information about the lenses and lensed stars obtained from the analysis
is summarized as follows.
\begin{enumerate}
\item
For all newly analyzed events, we measure the linear limb-darkening
coefficients of the surface brightness profile of the source stars.
\item
For all events with available CMDs of field stars, we measure the
Einstein radii and the lens-source proper motions. Among them, 5 events
(OGLE-2004-BLG-254, MOA-2007-BLG-176, MOA-2007-BLG-233/OGLE-2007-BLG-302,
MOA-2011-BLG-093, and MOA-2011-BLG-274) are found to have Einstein radii
less than 0.2 mas, making the lenses of the events candidates of very
low-mass stars or brown dwarfs.
\item The measured time scale $t_{\rm E} \sim 2.7$ days combined with
the small Einstein radius of $\sim 0.08$ mas of the event MOA-2011-BLG-274
suggests the possibility that the lens is a free-floating planet.
\item
For MOA-2009-BLG-174, we additionally measure the lens parallax and thus
uniquely determine the physical parameters of the lens. The measured
lens mass of $\sim 0.8\ M_\odot$ is consistent with that of a star blended
with the source, suggesting the possibility that the blend comes from the lens.
\item
We find no statistically significant planetary signals for any of the
events analyzed in this work. However, it is still possible to place
constraint on the range of the planetary separation and mass ratio.
For this purpose, we provide exclusion diagrams showing the confidence
levels of excluding the existence of a planet as a function of the
separation and mass ratio.
\end{enumerate}
\acknowledgments
Work by CH was supported by Creative Research Initiative Program
(2009-0081561) of National Research Foundation of Korea.
The MOA experiment was supported by JSPS17340074, JSPS18253002,
JSPS20340052, JSPS22403003, and JSPS23340064.
The OGLE project has received funding from the European Research
Council under the European Community's Seventh Framework Programme
(FP7/2007-2013) / ERC grant agreement no. 246678.
Work by BSG and AG was supported in part by NSF grant AST-1103471.
Work by BSG, AG, RWP, and JCY supported in part by NASA grant NNX08AF40G.
Work by JCY was supported by a National Science Foundation Graduate
Research Fellowship under Grant No.\ 2009068160.
CBH acknowledges the support of the NSF Graduate Research Fellowship
\#2011082275
TS was supported by the grants JSPS18749004, MEXT19015005, and JSPS20740104.
FF, DR and JS were supported by the Communaut{\'e}
fran\c{c}aise de Belgique - Actions de recherche concert{\'e}es -
Acad{\'e}mie universitaire Wallonie-Europe.
|
\section{Introduction}
The discovery of one or more pulsars in the inner parsecs around
Sgr~A*, the massive black hole (MBH) at the center of our
Galaxy, would provide an invaluable tool for studying the innermost
regions of the Galactic center (GC). Most of the current understanding
of the inner parsec comes from infrared observations of the nuclear
star cluster \citep[for a recent review, see][]{geg10}. The nuclear
star cluster is centered on Sgr~A*\ and consists of young massive stars
at a projected radius of $r\approx0.5$~pc and a dense collection of
B-stars (the ``S-stars'') within $r\leq0.04$~pc with the closest
orbit passing just $6\times10^{-4}$~pc (${\approx}100$~AU) from Sgr~A*\
\citep{sog02, gdm03}. Two decades of monitoring the orbits of these S-stars
has yielded the mass of the central object to be $M=4\times10^6M_{\odot}$,
unambiguously classifying it as a MBH \citep{gsw08, get09}.
Despite the success of tracking stellar orbits in the infrared, the
sensitivity of this method is ultimately limited by source confusion.
The detection of a radio pulsar at a similar distance with an orbital
period of $P_{orb}\lesssim100$~yr would provide unparalleled tests of
gravity in the strong-field regime. The timing of such a pulsar
could allow the measurement of the spin or quadrupole moment of the
MBH \citep{pl04, lw97, wk99, lwk12}. Additionally, a pulsar found anywhere in the inner few parsecs
of the Galaxy would provide a useful probe of the GC environment.
The mere detection of a pulsar would place constraints
on the star formation history (SFH) and measurements of
the dispersion measure and pulse broadening times would provide
information on the electron density distribution of the region.
However, even with the detection of almost 2000 radio pulsars in the Galaxy
\citep{PSRCAT} and
several directed searches of the GC, only five pulsars have been found within 15{\arcmin}
of Sgr~A*\ and the closest of these is $11^{\prime}$ away \citep{dcl09, jkl06, bjl11}.
While these few objects indicate the existence of a GC pulsar population,
the perceived dearth of pulsars near Sgr~A* is the result of
interstellar scattering from turbulent plasma,
which temporally broadens pulses to approximately $2000\nu_{\mbox{\scriptsize GHz}}^{-4}$~s
(where $\nu_{\mbox{\scriptsize GHz}}$ is the observing frequency in GHz) at the center of the Galaxy \citep{cl02}. Pulse broadening makes it almost impossible to detect even
long-period pulsars in periodicity searches at commonly used frequencies
($\nu\sim 1$~GHz).
To mitigate the deleterious effects of interstellar scattering,
periodicity searches of the GC have migrated to higher frequencies
($\nu \sim 10~\mbox{GHz}$). However, since pulsars have power-law
spectra of the form $S(\nu)\propto\nu^{\alpha}$ (with $\alpha<0$),
increasing the observing frequency also decreases the observable flux density.
To date, high-frequency searches have produced no new detections
using existing 100~m class telescopes \citep{deneva10, mkfr10}.
Even though the absence of pulsar detections in the central parsecs of the
GC is well explained by scattering effects, the existence of a GC pulsar
population was established by \citet{dcl09} based on the five pulsars on the outskirts of the region that cannot be explained as foreground disk objects.
Since future surveys can benefit from better knowledge of the pulsar populations in the GC,
we use a suite of multiwavelength observations to set constraints on the number and distribution of pulsars in the inner regions of the Galaxy on ${\sim}100$~pc and ${\sim}1$~pc scales. An illustration of the structure of the GC on these scales is shown in Figure \ref{fig:beam}.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.5\textwidth]{f1_color.eps}
\caption{Views of the inner GC region at radio, X-ray and infrared wavelengths. From top to bottom, the GC is shown in radio at 8.5~GHz as observed with the Green Bank Telescope (data courtesy Casey Law), in 0.5--7~keV X-rays as observed with the \emph{Chandra} ACIS-I instrument, and at J band ($1.25~\mu$m) as observed in 2MASS \citep{2MASS}. In each panel, a circle with radius {25\arcsec} (corresponding to 1~pc at 8.5~kpc) is centered on {Sgr~A*} with J2000 coordinates of ($17\h45\m40\fs0409$, $-29\degr00\arcmin28\farcs118$) given by \citet{rb04}.}
\label{fig:beam}
\end{center}
\end{figure}
In this paper, we present observational constraints on the pulsar populations in the GC.
A brief overview of the conventions and notations used in this paper are presented in Section~\ref{notation}.
In Section~\ref{radio_surveys}, population limits are set from the detections of pulsars in the inner 15{\arcmin} and from non-detections in the vicinity of {Sgr~A*} on parsec scales.
In Section~\ref{point_sources1}, a catalog of steep-spectrum radio sources in the inner 150~pc is considered.
Interferometric measurements of the spectrum of {Sgr~A*} are used in Section~\ref{diffuse_radio} to set upper limits on the pulsar population in the GC on arcsecond scales.
In Section~\ref{diffuse_gamma}, \emph{Fermi} observations of the diffuse gamma-ray flux of the inner degree of the GC are used to estimate the millisecond pulsar population in the GC.
In Section~\ref{massive_stars}, infrared observations of young massive stars are used to estimate the number of neutron stars produced in the inner parsec of the Galaxy.
\emph{Chandra} X-ray observations of pulsar wind nebulae are used to constrain the pulsar population in the inner 20~pc in Section~\ref{PWN_sec}.
In Section~\ref{supernova}, limits are set on the intrinsic neutron star population in the GC based on the estimated supernova rate.
Finally, in Section~\ref{discussion}, the estimates are summarized and discussed.
\section{Conventions and Notations}
\label{notation}
We note briefly a few conventions and notations that we adopt in this paper. The name ``Sgr~A*'' is used to describe both the MBH and the compact radio source at the dynamical center of the Galaxy. In the event where a distinction must be made (for example, Section~\ref{diffuse_radio}), ``Sgr~A*'' is taken to mean the observed radio source. All distances from Sgr~A* are given as a projected distance unless explicitly stated otherwise. The distance from Earth to Sgr~A*\ is taken to be $d = 8.5~\mbox{kpc}$ \citep{gsw08, get09}.
Pulse broadening times are taken from the NE2001 electron density model of \citet{cl02}. The NE2001 model accounts for the geometry of the scattering region in the GC, and as such, it is preferable over empirical fits to pulsars in the Galactic disk (e.g. \citealt{bcc04}), which do not consider the particular scattering geometry of the GC. As discussed in \citet{dcl09}, NE2001 tends to overestimate the scattering times for the five pulsars closest to Sgr~A*\ by factors of ${\sim}10^2-10^3$. However, these pulsars likely lie along the edges of the GC scattering region where slight (${\sim}0.1$~kpc) changes to the line of sight distance can cause dramatic (factors of ${\sim}10^4$) changes in scattering times with only modest (factor of ${\sim}2$) changes in dispersion measure. Additionally, we note that the scattering times of pulsars deeper in the GC will be highly constrained by the measured angular scattering of Sgr~A*\ itself.
Pulse broadening times are found at arbitrary observing frequencies by scaling the 1~GHz values given by NE2001 as $\propto\nu^{-4}$ \citep{lr99}. However, see \citet{lkm01} for potentially significant deviations from this scaling in highly scattered pulsars.
All pulsar population estimates are given as the number of active radio pulsars beamed toward the Earth, where a beaming fraction of $f_b = 0.2$ is assumed for all pulsars. The fixed beaming fraction of $f_b=0.2$ comes from a simple model in which the magnetic dipole moment is oriented randomly with respect to the rotation axis \citep{ec89}. A better empirical fit to the data is provided by the period-dependent model of \citet{tm98}, which typically finds $f_b\sim0.1$ for isolated pulsars. However, since the \citet{tm98} model does not include millisecond pulsars that may have beaming fractions as high as $f_b\approx0.5-0.9$ \citep{kxl98}, we adopt the $f_b=0.2$ value as a population-wide representative value. As this factor only shows up in our analysis as a multiplicative constant, it is trivial to scale our results to different beaming fractions.
In many of the limits presented below, it will be necessary to utilize a distribution for pulsar pseudo-luminosities ($L=Sd^2$). We adopt the power-law distribution of \citet{parkesVI2006} with a minimum cutoff in preference to distributions that do not require cutoffs like the log-normal model of \citet{fgk06}. Both are two parameter models and the power-law distribution provides a much better empirical fit to the observations \citep[see, e.g., Figure 6 of][]{parkesVI2006}. However, if the log-normal distribution were used instead of the power-law in the calculations below, the total pulsar populations predicted would be larger by factors of 10--100 owing to the much larger fraction of low luminosity objects. As a result, our adoption of the power-law pseudo-luminosity distribution is a conservative one in the sense that using another distribution would predict a \emph{larger} pulsar population.
Lastly, we note that in some cases our constraints will involve all types of pulsars, while others involve one of two subsets of pulsars: the ``canonical'' pulsars (CPs) and recycled or ``millisecond'' pulsars (MSPs).
CPs have periods of $P\sim1~\mbox{s}$, surface magnetic fields
$B\sim10^{12}~\mbox{G}$ and active radio lifetimes of
$\tau\sim10^7~\mbox{years}$,
while MSPs have periods $P\lesssim10~\mbox{ms}$, low surface magnetic fields $B\lesssim10^9~\mbox{G}$ and active radio lifetimes of
$\tau\sim 10^9-10^{10}~\mbox{years}$.
\section{Constraints from Pulsar Surveys of the GC}
\label{radio_surveys}
Several pulsar searches have been conducted in the inner degree of the Galaxy. To date, only five pulsars have been detected within $15^\prime$ of Sgr~A*, with none closer than $11^\prime$ \citep{PSRCAT}.
To have any chance of making detections in the inner few arcminutes of the Galaxy, higher observing frequencies must be used to overcome the roughly $2000\nu_{\mbox{\scriptsize GHz}}^{-4}$~s broadening times caused by scattering. Deep searches have been attempted at frequencies from 4 to 15~GHz, but have made no detections
in the inner few parsecs around {Sgr~A*}
\citep{jkl06,deneva10,mkfr10}.
Both the pulsar detections in ``low-frequency'' (2--3 GHz) surveys and the absence of detections in high-frequency (4--15 GHz) directed searches can be used to constrain the GC pulsar population.
\subsection{GC Pulsar Detections in Low-Frequency Surveys}
\label{ssec:lowfreq}
The five known pulsars within $15^\prime$ of {Sgr~A*} currently provide the best direct evidence for an intrinsic GC pulsar population. Of these five pulsars, two were detected at 3.1~GHz with the Parkes radio telescope \citep{jkl06} and three\footnote{One of these pulsars, J1746$-$2850, was independently discovered by \citet{bjl11} in a Parkes 6.5~GHz multibeam survey.} were detected at 2~GHz with the Green Bank Telescope \citep{dcl09}. In each of these surveys, the expected number of detectable disk pulsars in the field of view is $\ll1$. Thus, the detections strongly suggest a pulsar population in the GC that is distinct from that of the disk.
In an attempt to constrain the number and spatial distribution of the GC pulsars, \citet{dcl09} simulated the pulsar population to determine what would be consistent with the survey detections. A simple two-component density model of the form
\begin{equation}
\label{ellipsoidal}
n_{GC}\propto\exp\left(-\frac{h^2}{H_{GC}^2}\right)\exp\left(-\frac{r^2}{R_{GC}^2}\right)
\end{equation}
was adopted with $H_{GC}=26$~pc fixed to coincide with the scale height of the scattering screen in the NE2001 electron density model of \citet{cl02}. The density distribution was normalized so that, for a given $R_{GC}$, there were a total of $N_{GC}$ pulsars (not necessarily beamed toward Earth) associated with the population.
Using Monte Carlo methods, \citet{dcl09} then generated 1000 pulsar populations consistent with Equation~\ref{ellipsoidal} for each $(R_{GC}, N_{GC})$ pair and determined how many of these pulsars would have been detected in their 2.1~GHz survey. Searching over a grid of values and performing a maximum likelihood analysis, \citet{dcl09} found lower bounds of $N_{GC}\gtrsim2000$ and $R_{GC}\gtrsim0.3$~kpc for the parameters. Thus, this analysis provides additional evidence for an intrinsic pulsar population in the GC.
Although \citet{dcl09} have set a lower bound on $N_{GC}$, this parameter applies to the entire GC population and does not necessarily translate into a lower bound on the number of pulsars within a particular distance from {Sgr~A*}. For example, a distribution with $N_{GC}=2000$ and $R_{GC}=0.3$~kpc will produce a very different number of pulsars within 100~pc of {Sgr~A*} than will a distribution with $N_{GC}=2000$ and $R_{GC}=0.6$~kpc. Regardless, the existence of a pulsar population in the GC is firmly established.
We note briefly that the primary reason the \citet{dcl09} analysis does not find upper bounds on the parameters $N_{GC}$ and $R_{GC}$ even after extending the grid to $N_{GC}=10^4$ and $R_{GC}=5$~kpc is that the only constraints come from the detections in the survey region ($r\lesssim50$~pc). Incorporating survey results from the inner few degrees of the Galaxy would certainly introduce upper bounds to the parameters.
With this in mind, it is instructive to consider the results at a fixed $R_{GC}$. From Figure 4 of \citet{dcl09}, we see that a wide range of $N_{GC}$ values with $N_{GC}\gtrsim500$ are equally likely for $R_{GC}=0.1$~kpc. Thus, a very conservative lower bound on the number of pulsars in the inner 100~pc of the Galaxy that are beamed toward Earth is $N_{psr}\gtrsim100$ (where we have assumed a beaming fraction of 0.2).
\subsection{High-Frequency Pulsar Searches of the Central Parsec}
\label{gbt_search}
Recently, searches of the central few parsecs around Sgr~A*~have been conducted with the Green Bank Telescope (GBT) at 5~GHz and 9~GHz \citep{deneva10} and at 15~GHz \citep{mkfr10}.
No pulsars were detected in any of these searches. We follow a similar analysis to that of \citet{mkfr10} to estimate an upper limit to the pulsar population based on the absence of detections.
\subsubsection{Observations}
\label{sssec:obs}
The 5 and 9~GHz observations were carried out by \citet{deneva10} in 2006. Since no pulsar candidates were detected, limits on the flux density of periodic signals may be set using the radiometer equation
\begin{eqnarray}
\label{eq:flux_max}
S_{min,\nu} = \frac{mT_{sys}}{\eta G \sqrt{N_h N_{pol} \Delta\nu T_{obs}}},
\end{eqnarray}
where $T_{sys}$ is the system temperature, $G$ is the telescope gain, $\eta\approx0.8$ is a correction factor that accounts for system imperfections and the digitization of the signal, $N_h=16$ is the maximum number of harmonics summed in a periodicity search, $N_{pol}=2$ is the number of polarization channels summed, $\Delta\nu=800$~MHz is the receiver bandwidth and $T_{obs} = 6.5$~hr is the observation time. The telescope gain is given by $G=1.85~\mbox{K Jy}^{-1}$ and $G=1.8~\mbox{K Jy}^{-1}$ for observing frequencies of 5 and 9~GHz, respectively. The value of $m$ is determined by the detection significance threshold set at $m\sigma$. In this FFT search the threshold was set to $6\sigma$, so $m=6$.
The system temperature of the telescope is given by\footnote{We take $T_{rec}$ to include all non-astronomical contributions to the system temperature from the receiver, spillover effects, and the atmosphere. } $T_{sys} = T_{rec}+T_{bg}$.
The receiver temperature of the GBT\footnote{As provided in the GBT Proposer's Guide: \texttt{http://www.gb.nrao.edu/gbtprops/man/GBTpg.pdf}}
is 18~K at 5~GHz and 27~K at 9~GHz. The dominant contribution to the background temperature is the bright extended Sgr~A~Complex (comprised of Sgr~A~East and Sgr~A~West), which surrounds Sgr~A*.
We may set lower bounds on this background using data from a multiwavelength survey by \citet{lyc08}, which imaged the GC at 1.4, 5 and 9~GHz.
Only lower bounds may be set since a non-trivial iterative scheme was used to subtract out the noise contributions from the atmosphere.
\citet{lyc08} found the flux density of the Sgr~A Complex to be $85~\mbox{Jy beam}^{-1}$ at 5~GHz and $39~\mbox{Jy beam}^{-1}$ at 9~GHz, which translate to background temperatures of $T_{bg}=157$~K and $T_{bg}=70$~K, respectively. Thus, the system temperature is $T_{sys}=175$~K at 5~GHz and $T_{sys}=97$~K at 9~GHz. From Equation~\ref{eq:flux_max}, we find that $S_{min,\nu} = 29~\mu$Jy at 5~GHz and $S_{min,\nu} = 17~\mu$Jy at 9~GHz.
Likewise, \citet{mkfr10} conducted a search for pulsars at 14.8 and 14.4~GHz using the GBT in 2006 and 2008. The \citet{mkfr10} search used a $10\sigma$ detection threshold and combined observations to get an effective observation time of $T_{obs}\approx9.75$~hr. The system temperature was determined to be $T_{sys}\approx35$~K by firing a noise diode on a calibrator source. No pulsars were detected and a $10\sigma$ detection threshold flux density of $S_{min,\nu}\approx10~\mu$Jy is set from the 14.4~GHz measurements.
\subsubsection{Upper Limits on Observable Pulsar Population}
\label{sssec:hf_ul}
Assuming that a given pulsar in the GC will be detected with some probability $p_d$, binomial statistics can be used to find the maximum number of pulsars consistent with zero detections.
The simplest way to determine $p_d$ is to set it equal to the fraction of pulsars bright enough to be seen in each survey when placed at Sgr~A*. This fraction can be estimated from the 1.4~GHz pseudo-luminosity function, which is given by
$dN/d \log{L} \propto L^{-\beta}$
taken over a range of pseudo-luminosities from
$L_{min}=0.1~\mbox{mJy}~\mbox{kpc}^2$ to $L_{max} = 10^4~\mbox{mJy}~\mbox{kpc}^2$.
We adopt a value of $\beta=-0.7$ as an average of the two fits found by \citet{parkesVI2006} of over 1000 pulsars observed in the Parkes Multibeam Survey.
The detection probability is then given by $p_d=f_L(L>L_{det})$, where
\begin{eqnarray}
\label{eq:ldetect}
L_{det} = S_{min,\nu}\left(\frac{1.4~\mbox{GHz}}{\nu}\right)^{-1.7}d^2_{gc}
\end{eqnarray}
and $S_{min,\nu}$ is the minimum detectable flux density of the search.
Using the $S_{min,\nu}$ values found by each search (see Table~\ref{tab:hf_vals}) we find that $p_d=(0.027, 0.020, 0.015)$ at $\nu=(5,9,15)$~GHz.
Given the above detection probabilities and the lack of any detections in the surveys, the upper limits to the number of pulsars (at 99\% confidence level) are found to be $N=(170, 230, 299)$ for the 5, 9, and 15~GHz observations, respectively. However, since a pulsar with a spin period less than the pulse broadening time would have a greatly reduced chance of being detected, the calculated upper limits are for pulsars with $P\gtrsim\tau_{sc}$. Using the 1~GHz scattering time from the NE2001 model of \citet{cl02} and scaling ($\propto\nu^{-4}$) to the appropriate frequency, the scatter broadening times are found to be $\tau_{sc}=(4.2, 0.44, 0.05)$~s at observing frequencies of $\nu=(5,9,15)$~GHz.
\begin{table*}[ht]
\begin{center}
\begin{threeparttable}
\caption{Pulsar Upper Limits of High-Frequency Surveys of the Central Parsecs}
\label{tab:hf_vals}
\begin{tabular}{ccccccccc}
\toprule
\toprule
$\nu$ & $S_{min}$ & $\tau_{sc}$ & $f_L(L>L_{det})$ & $N(P\gtrsim\tau_{sc})$ & $f_{P}(P>\tau_{sc})$ & $N/f_P$ & $r$ & Ref\\
(GHz) & $(\mu$Jy) & (s) & & & & & (pc) & \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9)\\
\cmidrule{1-9}
4.85 & 29 & 4.2 & 0.027 & $<170$ & 0.06 & $<2830$ & 3.0 & (1)\\
6.6 & 550\tnote{$\dagger$} & 1.28 & 0.002 & $<2220$ & 0.17 & $<13000$ & 4.0 & (2) \\
8.4 & 200 & 0.46 & 0.003 & $<1380$ & 0.43 & $<3200$ & 3.0 & (3) \\
8.50 & 17 & 0.44 & 0.020 & $<230$ & 0.43 & $<535$ & 1.5 & (1)\\
14.4 & 10 & 0.05 & 0.015 & $<299$& 0.63 & $<475$ & 1.0 & (4)\\
\bottomrule
\end{tabular}
\begin{tablenotes}
\footnotesize
\item[] {\bf Notes.} Columns 1 and 2 give the center frequency and minimum detection threshold, respectively, of a given survey. Column 3 gives the scattering time at the survey center frequency of a pulse originating from the GC as calculated with the NE2001 model \citep{cl02}. Column 4 gives the fraction of pulsars bright enough to be detected at the GC with the survey sensitivity (see Equation~\ref{eq:ldetect}). Column 5 gives 99\% upper limit values to the number of pulsars that could potentially have been seen by the survey (that is, have $P\gtrsim\tau_{sc}$). Column 6 gives the fraction of pulsars in the ATNF catalog \citep{PSRCAT} that have periods longer than the scattering time and Column 7 gives an estimated upper bound for the total number of pulsars in the GC. Column 8 gives the projected radial distance of the beamsize at the distance of {Sgr~A*} and Column 9 gives the reference from which the survey was taken.
\item[] ${}^\dagger$The value provided for $S_{min}$ here is 9 times that given in Section~2.2 of \citet{bjl11}, which we believe to be a calculation error.
\item[] {\bf References.} (1) \citealt{deneva10}; (2) \citealt{bjl11}; (3) \citealt{jkl06}; (4) \citealt{mkfr10}
\end{tablenotes}
\end{threeparttable}
\end{center}
\end{table*}
\subsubsection{Implications for Total GC Pulsar Population}
The upper limits of Section~\ref{sssec:hf_ul} are only valid for pulsars within certain period ranges. To make an estimate of the total number of pulsars, knowledge of the underlying pulsar period distribution is needed. Though the period distribution of pulsars in the inner parsecs of the GC is entirely unknown, a reasonable approximation would be to assume the same distribution as the local pulsar population (to reduce observational biases). From the ATNF
pulsar catalog\footnote{\texttt{http://www.atnf.csiro.au/research/pulsar/psrcat/}}
\citep{PSRCAT}, we see that there are 88 pulsars within 1~kpc of Earth. Of these, 5 have $P>4.2$~s, 38 have $P>0.44$~s and 55 have $P>50$~ms. This gives $f_P(P>4.2~\mbox{s}) = 0.06$, $f_P(P>0.44~\mbox{s})=0.43$ and $f_P(P>50~\mbox{ms})=0.63$ for the fractions of pulsars with periods greater than the scatter broadening times.
Assuming these values are representative of the GC population, we can estimate upper bounds on the total number of pulsars (regardless of period) to be $N_{max}=N/f_P$. Applying these corrections to the estimates of Section~\ref{sssec:hf_ul} gives $N_{max} = (2830, 535, 475)$ for observing frequencies of $\nu=(5,9,15)$~GHz.
The half-power beam width of the GBT is $\theta(\nu)\approx150^{{}^{\prime\prime}}(\nu/5~\mbox{GHz})^{-1}$. At Sgr~A*, the projected radii of these beams are $r\approx(3.0, 1.5, 1.0)$~pc at 5, 9 and 15~GHz, respectively. Thus, we estimate that there are as many as $N_{max}<2830$ pulsars beamed toward Earth within $r\approx3.0$~pc of Sgr~A*, $N_{max}<535$ within $r\approx1.5$~pc and $N_{max}<475$ within $r\approx1$~pc.
The results of this section are summarized in Table \ref{tab:hf_vals}. In addition to the surveys of \citet{deneva10} and \citet{mkfr10}, we also include for reference the less sensitive GC pointings from surveys by \citet{jkl06} and \citet{bjl11}.
\subsubsection{Caveats}
A number of assumptions are made in the upper limit estimates of the previous sections, so it is important to consider what happens if the assumptions fail. The first assumption is that the probability of detecting a pulsar (with period large enough not to be smeared out by interstellar scattering) is equal to the fraction of pulsars with 1.4~GHz pseudo-luminosities large enough to be detected at the GC. Since the detection probability only considers the best-case sensitivity of the telescope and ignores any effects of radio-frequency interference at the telescope end or intermittency at the pulsar end, it is likely to be an overestimate. An overestimate of the detection probability would result in an underestimate of the upper bound on the number of pulsars.
Another assumption is that the 1.4~GHz pseudo-luminosity distribution of pulsars in the GC is the same as those in the Galactic field. However, if the GC region contains a larger number of low-luminosity MSPs, then the detection probability will again be overestimated and the upper limit underestimated. One could also imagine the case where enough bright young pulsars exist in the GC as a result of recent star formation to skew the pseudo-luminosity distribution to higher luminosities. In that case, the detection probability would be an underestimate and the upper limits an overestimate.
We have also assumed that the upper limits are for pulsars with $P>\tau_{sc}$. However, as the scattering time gets to be a significant fraction of the pulsar period, it will start to smear out the signal and reduce the number of detectable harmonics. Since this is a gradual process, it will likely make some pulsars undetectable even with $P>\tau_{sc}$ . As a result, we would be overestimating the minimal detectable period, which would cause the upper bound to be an underestimate.
Finally, we have assumed that the period distribution of the pulsars within 1~kpc of Earth is representative of the GC population. This certainly does not have to be the case, as the star formation histories of the Galactic field and GC are likely to be different. For example, the GC could potentially have a much higher concentration of young pulsars and old MSPs as compared to the Galactic disk. The increased stellar encounter rate in the GC could favor MSP production and any recent starburst would favor young pulsars. In both cases, the periods would be biased low. Thus, the period distribution would be skewed lower than assumed and the fraction of pulsars with periods greater than a certain value will be overestimated. This will result in the upper bound being underestimated.
Since most of the assumptions made tend to decrease the upper limits, our estimates are best interpreted as the most restrictive upper bounds to the pulsar population in the inner few parsecs of the GC.
\section{Radio Point Sources}
\label{point_sources1}
Motivated by the study of GC pulsar search methods by \citet{cl97}, \citet{lc08} performed a VLA survey of compact radio sources in the inner degree of the Galaxy. Though pulsars cannot be identified by their pulsed emission in an imaging survey, promising pulsar candidates may be found by looking for steep-spectrum sources with angular diameters consistent with the angular broadening of point sources caused by scattering at locations near {Sgr~A*} (${\approx}1\arcsec$~at 1~GHz). Of the 170 compact radio sources cataloged, \citet{lc08} estimate that the number of pulsars included is of order ${\sim}10$. Based on this survey, upper limits to the pulsar population within $1^\circ$ (${\approx}150$~pc) of {Sgr~A*} may be estimated.
\subsection{Observations}
The survey was conducted at observing frequencies of 1.4 and 5~GHz with the VLA in the A~configuration. A total of 13 fields arranged in a hexagonal grid covered the region of the GC out to roughly $1^\circ$ (150~pc) from {Sgr~A*} (the half-power radius of the VLA primary beam is $15\arcmin$ at 1.4~GHz). The typical resolution for the survey was a synthesized beam size of $2\farcs4\times1\farcs3$.
Sources were identified using a method similar to that of \citet{lc98}. Essentially, a histogram of intensities was constructed from the image of the primary beam for each field. If the field just contained noise, the intensity histogram would be a Gaussian with a mean of zero and a standard deviation equal to the thermal noise of 0.05~mJy per synthesized beam. Sources could then be determined by looking for deviations from this noise-only histogram. In practice, the histogram was found to have larger tails than a Gaussian, with zero mean and a standard deviation of ${\approx}0.5~\mbox{mJy beam}^{-1}$. Since the resolution is comparable to the scattering size of a point source at the distance of {Sgr~A*}, a $10\sigma$ detection threshold of $S_{det}\approx5$~mJy was adopted for the survey.
\subsection{Pulsar Population Estimate}
Given that $N_{obs}\sim10$ pulsars were likely observed, the total pulsar population in the survey region can be estimated as $N_{psr} \sim N_{obs}/f_L$, where $f_L$ is the fraction of pulsars luminous enough to be detected at the distance of {Sgr~A*}. Taking the survey detection threshold to be $S_{det}=5$~mJy at 1.4~GHz, a pulsar must have a 1.4~GHz pseudo-luminosity of at least $L_{det}=360~\mbox{mJy kpc}^2$ to be detected at the distance of {Sgr~A*}.
The fraction of pulsars with $L>L_{det}$ can be determined from the 1.4~GHz pulsar luminosity function, which has the form
$dN/d \log{L} \propto L^{-\beta}$.
The range of pulsar pseudo-luminosities is taken from
$L_{min} = 0.1~\mbox{mJy}~\mbox{kpc}^2$ to $L_{max} = 10^4~\mbox{mJy}~\mbox{kpc}^2$ and the exponent in the distribution function is taken to be $\beta = 0.7$ \citep{parkesVI2006}. From this distribution, the fraction of pulsars luminous enough to be detected is $f_L=3\times10^{-3}$. For $N_{obs}\sim10$ pulsars detected in the survey, we expect a total population of $N_{psr}\sim3000$ pulsars within $1^\circ$ ($150$~pc) of {Sgr~A*}.
Though $N_{psr}\sim3000$ is the nominal population estimate from the survey, a broader range results if the assumptions do not exactly hold. For instance, \citet{lc08} estimate that $N_{obs}\sim10$ of the unidentified steep-spectrum point sources will ultimately turn out to be radio pulsars. However, this number could range from zero to about 30. If one takes $N_{obs}=30$, repeating the above analysis gives a pulsar population of $N_{psr}\sim10^4$. Additionally, one may consider the case in which no pulsars were detected. Despite a follow-up observation of 15 of the pulsar candidates in this survey by \citet{deneva10} with the GBT, none of the candidates have to date been confirmed. Assuming zero detections, an analysis similar to that in Section~\ref{gbt_search} gives an upper limit to the pulsar population of $N_{psr}\leq1500$ at a 99\% confidence level. These two extremes illustrate that although the survey allows for an estimate of the pulsar population in the inner degree of about 3000, the actual number could be below 1500 or as high as $10^4$. As a result, we take $N_{psr}\lesssim10^4$ as a conservative upper bound.
Finally, we note that although the survey covers the region within $1^\circ$ ($150$~pc) of {Sgr~A*}, there will be reduced sensitivity in the field centered on {Sgr~A*}. The reduced sensitivity is the result of increased background temperatures and greater sidelobes from the extended structure of the inner GC. In addition, since the scatter-broadening of point sources in the vicinity of {Sgr~A*} (${\approx}1\arcsec$) is comparable to the resolution of the survey, source confusion may become important in the innermost regions of the GC. Therefore, this survey would be largely insensitive to a fairly compact population of pulsars in the inner tens of arcseconds around {Sgr~A*}.
\section{Radio Spectrum of Sgr A*}
\label{diffuse_radio}
In this section, limits are placed on the maximum allowable number of
pulsars in the inner parsecs of the GC based on radio interferometer
observations of {Sgr~A*} on arcsecond scales ($1\arcsec\approx0.04$~pc
at 8.5~kpc). Due to the finite resolution of interferometers
and the broadening of angular diameters as a result of the interstellar
scattering of radio waves, the {Sgr~A*} radio source is actually
extended (${\approx}1\arcsec$ at 1~GHz). Flux measurements
of {Sgr~A*} will therefore include a contribution from a collection of
pulsars, if such a population exists. Although these pulsars will
be unresolved, upper limits on the total population may be set based
on the total flux density of {Sgr~A*} in a manner analogous
to similar constraints placed on pulsars in globular clusters \citep{fg90}.
We consider a model in which the observed flux density of {Sgr~A*} is
actually the combination of two components. The first component is
that due to radio emission from the immediate environment
of the MBH itself, which we assume
is described accurately by high frequency observations where the pulsar component is negligible.
The second component is that due to the population of
pulsars near the MBH. This pulsar component becomes
important at lower frequencies both because radio pulsars
typically have steep spectra
and the angular resolution of radio telescopes scales with frequency such
that a larger region around the MBH is sampled at
lower frequencies. By requiring that this model flux be consistent with
existing observations, constraints may be set on the maximum number of
pulsars allowed in the inner parsec of the Galaxy.
\subsection{Observations}
Two different measurements of the spectrum of Sgr~A* over a wide
range of frequencies are considered \citep{an05, falcke98}. \citet{an05} conducted simultaneous measurements of Sgr~A* from 300~MHz to 43~GHz using the VLA (A-configuration) and the GMRT. \citet{falcke98} made simultaneous measurements of Sgr~A* using the VLA (A-configuration), the Berkeley-Illinois-Maryland Array, the Nobeyama 45~m telescope, and the Institut de Radioastronomie Millimetrique (IRAM) 30~m telescope from 1.4~GHz to 235~GHz. Both groups observed a broken power-law spectrum with a break around 10~GHz. Since the power-law spectrum of pulsars decreases with increasing frequency, a population of pulsars
contributes significantly only at lower frequencies. As a result, we consider
the spectrum of {Sgr~A*} only below the break frequency at 10~GHz.
\subsection{Spectral Model}
\label{ss_model}
We model the measured flux density of the compact radio source Sgr~A*\ as the sum
of contributions from a collection of pulsars and a point source associated with the MBH attenuated by free-free absorption according to
\begin{eqnarray}
S_{\mbox{\scriptsize Sgr}}(\nu) &=&\left[ S_{\mbox{\scriptsize BH,$\nu_0$}} \left(\frac{\nu}{\nu_0}\right)^{\alpha_{\mbox{\scriptsize bh}}} +N(\nu)S_{\mbox{\scriptsize psr,$\nu_0$}} \left(\frac{\nu}{\nu_0}\right)^{\alpha_{\mbox{\scriptsize psr}}} \right] \nonumber \\
&& \times\exp( -\nu_{f}^2/\nu^2).
\label{eq:spectrum}
\end{eqnarray}
The emission from the immediate vicinity of the MBH, $S_{\mbox{\scriptsize BH,$\nu_0$}}$,
is taken to be a point source with a power-law spectrum
with spectral index $\alpha_{\mbox{\scriptsize bh}}$.
$N(\nu)$ is the number of pulsars contained in the
solid angle of the effective point-spread function or beam size,
which depends strongly on frequency (see Section~\ref{ssec:ang_res}).
We use a simplified scaling for the free-free absorption that ignores the frequency
dependence of the Gaunt factor.
The free-free absorption factor has a turnover frequency
$\nu_f$ \citep{an05}. The flux density per pulsar, $S_{\mbox{\scriptsize psr,$\nu_0$}}$, is taken to
be the mean of the 1.4~GHz pulsar pseudo-luminosity distribution given by $dN/d \log{L}
\propto L^{-0.7}$ \citep{parkesVI2006} with a lower cutoff of $L_{min} = 0.1 \mbox{ mJy kpc}^2$ and an upper cutoff of $L_{max} = 10^4 \mbox{ mJy kpc}^2$.
The mean observed
flux density at 1.4~GHz is $S_{\mbox{\scriptsize psr,$1.4$}} = 99~\mu\mbox{Jy}$.
Since pulsar radio flux scales
with frequency as a simple power-law, we can scale our flux density as
$S_{\nu} \propto \nu^{\alpha_{\mbox{\scriptsize psr}}}$. We fix
$\alpha_{\mbox{\scriptsize psr}} = -1.7$ as a nominal value for the pulsar spectral index
\citep{mkk00, lyl95}, but also consider values of $-1.0$ and $-2.5$ to test any major spectral index
dependence.
\subsection{Effective Angular Resolution}
\label{ssec:ang_res}
The number of pulsars included in a flux measurement of Sgr~A*\ can be written
as an integral of the number of pulsars per unit solid angle,
\begin{eqnarray}
N(\nu) = \int d\Omega \frac{dn_p}{d\Omega}.
\end{eqnarray}
The integral is over the effective solid angle
$\Omega_{\rm eff}(\nu)= (\pi/4)\theta_{\rm eff}^2(\nu)$, where
\begin{eqnarray}
\theta_{\rm eff}(\nu)
= \left[\theta_{\rm b}^2(\nu) + \theta_{\rm sc}^2(\nu) \right]^{1/2}
= \left(\theta_{\rm b_0}^2\nu^{-2} + \theta_{\rm sc_0}^2\nu^{-4} \right)^{1/2}
\end{eqnarray}
is the effective resolution with the subscript ``0'' representing values at 1~GHz and the frequencies are
in GHz units. The effective resolution is the quadrature sum of the resolution of the interferometer and
the angular extent of Sgr~A*\ caused by scattering.
The scaling for the synthesized array beam $\theta_{\rm b}$ $\left(\propto\nu^{-1}\right)$ assumes a
fixed array configuration and the scattering diameter $\theta_{\rm sc}$ $\left(\propto\nu^{-2}\right)$ scales
in conformance to
measurements of Sgr~A*\ and OH/IR masers \citep[e.g.,][]{fcc94}.
Our treatment assumes that scattered images are circular whereas in fact some
are elliptical, but given that we are making order of magnitude estimates of
pulsar numbers, the differences are not important.
The angular diameter of Sgr~A*\ is dominated by interstellar
scattering at low frequencies, with an observed major axis of
$\theta_{\rm sc_0} = 1\farcs2$ \citep{bgf06}.
For the VLA in the A configuration and ignoring any effects of foreshortening, the half-power beam width is $\theta_{\rm b_0} = 1\farcs95$ \citep{bridle89}.
The effective resolution of the VLA observations is then
\begin{equation}
\theta_{\rm eff} (\nu) =
1\farcs2\, \nu^{-2}\left(1 + 2.6~\nu^{2}\right)^{1/2}. \end{equation}
The two contributions are equal
at $\nu = 0.6$~GHz, so at frequencies lower than this the resolution
is completely scattering dominated and
the resolution solid angle scales steeply with frequency as $\nu^{-4}$.
Additionally, a single data point measured with the GMRT will be considered in our analysis, so a similar effective resolution must be constructed for this telescope. \citet{royrao04} measure the resolution to be $11\farcs4\times7\farcs6$ at 620~MHz. Converting this ellipse to a circle of equal area and scaling to 1~GHz gives $\theta_{\rm b_0} = 5\farcs77$ for the GMRT. Combining this with scattering as above, gives
\begin{equation}
\theta_{\rm eff} (\nu) =
1\farcs2\, \nu^{-2}\left(1 + 23.1~\nu^{2}\right)^{1/2}. \end{equation}
For the GMRT, the two components of the effective resolution are equal at $\nu=0.2$~GHz.
\subsection{Candidate Pulsar Distributions}
\label{ssec:cand_dist}
We choose three physically-motivated distributions as model
pulsar populations. The distributions are illustrated below in Figure \ref{fig:densities}.
The first (Model A) assumes a constant number of pulsars per unit solid angle,
$dn_p/d\Omega = $~constant,
so the number of pulsars scales as
\begin{eqnarray}
N_A(\nu) = N_1 \left[\frac{\Omega_{\rm eff}(\nu)}{\Omega_1}\right],
\end{eqnarray}
where $\Omega_1$ is the solid angle enclosing $N_1$ pulsars. In the fitting below, $\Omega_1$ is set so that $N_1$ gives the number of pulsars in the inner parsec.
Referring to Equation~\ref{eq:spectrum}, it may be seen that when scattering dominates the effective
resolution, the contribution
to the unabsorbed spectrum from pulsars
increases very rapidly as $N_A(\nu) \nu^{\alpha_{\rm psr}} \propto \nu^{-5.7}$.
Free--free absorption attenuates much of the flux, thus allowing a
significant pulsar population to remain hidden in spectral measurements.
In Model B, we assume that $dn_p/d\Omega \propto \Omega^{-0.7}$,
corresponding to the surface density scaling
observed for Wolf-Rayet and O-star populations in the
inner parsec \citep{geg10}. This yields
\begin{eqnarray}
N_B(\nu) = N_1 \left[\frac{\Omega_{\rm eff}^{0.3}(\nu)}{\Omega_1^{0.3}} \right]. \end{eqnarray}
However, the observed populations of Wolf-Rayet and O-stars have an inner cutoff at $\theta\approx1^{\prime\prime}$ \citep{bartko10}. This core may affect the pulsar population in many ways, but we shall just consider two here. In both distributions, the pulsar surface density goes as $dn_p/d\Omega \propto \Omega^{-0.7}$ outside the inner cutoff, as before. Inside the cutoff, one of the distributions (call it Model B-1) has $dn_p/d\Omega = \mbox{const}$ and the other (call it Model B-2) has $dn_p/d\Omega = 0$. These models give
\begin{eqnarray}
N_{B1}(\nu) = \left\{
\begin{array}{l l}
N_1 \left[\frac{\displaystyle\Omega_{\rm eff}^{0.3}(\nu) -0.7\Omega_0^{0.3}}
{\displaystyle\Omega_1^{0.3}-0.7\Omega_0^{0.3}} \right], & \quad \Omega_{\rm eff}\geq\Omega_0\\
& \\
N_1 \left[\frac{\displaystyle0.3\Omega^{0.7}_0\Omega}
{\displaystyle\Omega_1^{0.3}-0.7\Omega_0^{0.3}} \right]& \quad \Omega_{\rm eff}<\Omega_0\\
\end{array} \right.
\end{eqnarray}
and
\begin{eqnarray}
N_{B2}(\nu) = \left\{
\begin{array}{l l}
N_1 \left[\frac{\displaystyle\Omega_{\rm eff}^{0.3}(\nu) -\Omega_0^{0.3}}
{\displaystyle\Omega_1^{0.3}-\Omega_0^{0.3}} \right], & \quad \Omega_{\rm eff}\geq\Omega_0\\
0 & \quad \Omega_{\rm eff}<\Omega_0\\
\end{array} \right.
\end{eqnarray}
pulsars enclosed within $\Omega_{\rm eff}$.
In Model C, we consider a compact population of pulsars contained in a solid
angle much smaller than any resolution solid angle so that
$dn_p/d\Omega$ is effectively a delta function.
A compact distribution close to Sgr~A*\ could conceivably arise as
a product of dynamical friction \citep{morris93, meg00}.
Here we simply have
\begin{eqnarray}
N_C(\nu) = N_1.
\end{eqnarray}
In addition to the above three, one may consider other models for the pulsar distribution. For example, the pulsars could be arranged in a central core with a diffuse halo. However, most of these other distributions can be made as combinations of those we consider. As a result, we do not expect the final answers to change by more than an order of magnitude.
\begin{figure}[h!]
\begin{center}
\includegraphics[width=0.5\textwidth]{f2.eps}
\caption{Pulsar number density as a function of angular separation from Sgr~A*~for each of the three distribution models. The distributions have been normalized in this figure so that each model produces the same number of pulsars enclosed within the inner parsec ($\theta_r \approx 25^{\prime\prime}$). See Section~\ref{ssec:cand_dist} for a description of each model.}
\label{fig:densities}
\end{center}
\end{figure}
\subsection{Model Fitting}
To test our distributions, we calculated the $\chi^2$ values of the model given by Equation~\ref{eq:spectrum} with each of the three distributions against the VLA and GMRT data from \citet{an05} and \citet{falcke98}.
We fix the pulsar spectral index at $\alpha_{\mbox{\scriptsize psr}} = -1.7$ and the flux density per pulsar at $S_{\mbox{\scriptsize psr,$\nu_0$}} = 99~\mu \mbox{Jy}$ at 1.4~GHz, as described in Section~\ref{ss_model}.
The flux from the point source associated with the MBH is normalized such that the total measured flux of Sgr~A*\ exactly matches the data at 8.45~GHz.
The spectral index of the MBH point source ($\alpha_{\mbox{\scriptsize bh}}$), the number of pulsars ($N_1$) and the free-free cutoff frequency ($\nu_f$) are allowed to vary.
The number of pulsars, $N_1$, is taken within an angular distance of $\theta = 25\arcsec$, which corresponds to a projected radial distance of $r\approx1$~pc from Sgr~A*.
The allowed ranges for each parameter were chosen to be consistent with current measurements and are presented below in Table~\ref{params}.
For reference, the best fit values for the model with no pulsars present are $\alpha_{\mbox{\scriptsize bh}}=0.15$ and $\nu_f=0.26$~GHz.
For each grid point in the three-dimensional parameter space, we
calculate $\chi^2$ between the measured flux and our model and evaluate the likelihood function assuming independent Gaussian
statistics for measurement errors,
\begin{eqnarray}
\mathcal{L}(N_1, \alpha_{\mbox{\scriptsize bh}}, \nu_f) &=& \prod_{i=1}^{N_p} \left({2\pi\sigma_i^2}\right)^{-\frac{1}{2}} \exp \left\{ -\frac{\left[S_{\mbox{\scriptsize Sgr}}(\nu_i) - S_{obs}(\nu_i)\right]^2}{2\sigma_i^2} \right\} \nonumber \\
&\propto& \exp \left( -\frac{1}{2} \chi^2 \right).
\end{eqnarray}
The likelihood function is then marginalized over $\alpha_{\mbox{\scriptsize bh}}$ and $\nu_f$ to get a distribution for $N_1$. The marginalized likelihood functions for the number of pulsars within 1 pc of Sgr~A* are plotted in Figure~\ref{fig:likelihoods}.
\resizebox{0.5\textwidth}{!}{
\begin{threeparttable}
\caption{Searched Parameters for Each Pulsar Distribution Model}
\label{params}
\centering
\begin{tabular}{cccc}
\toprule
\toprule
& \multicolumn{3}{c}{Parameter Ranges\tmark} \\
\cmidrule(l){2-4}
Model & $N_1$ & $\alpha_{\mbox{\scriptsize bh}}$ & $\nu_f~\mbox{(GHz)}$\\
\midrule
A & [ 0, 50000, 100 ] & [ 0.05, 0.40, 0.01 ] & [ 0.05, 1.00, 0.01 ] \\
B & [ 0, 50000, 100] & [ 0.05, 0.40, 0.01 ] & [ 0.05, 1.00, 0.01 ] \\
C & [ 0, 5000, 10 ] & [ 0.05, 0.40, 0.01 ] & [ 0.05, 1.00, 0.01 ]\\
\bottomrule
\end{tabular}
\begin{tablenotes}
\footnotesize
\item[a] Data given as [ min, max, step size ]
\end{tablenotes}
\end{threeparttable}}
\begin{figure}[h!]
\centering
\includegraphics[width=0.5\textwidth]{f3.eps}
\caption{Likelihood functions for three pulsar distribution models (see Section~\ref{ssec:cand_dist} for a description of each model). The likelihoods have been marginalized over the parameters $\alpha_{\mbox{\scriptsize bh}}$ and $\nu_{f}$ and normalized so that
$\int \mathcal{L}(N)=1$.
In each case, the pulsar spectral index was taken to be $\alpha_{\mbox{\scriptsize psr}}=-1.7$.}
\label{fig:likelihoods}
\end{figure}
\subsection{Results}
The best-fit parameters for the maximum likelihood (ML) number of pulsars within the inner parsec of the GC for each model is provided in Table \ref{tab:model_params}
and the resulting spectra are plotted in Figure \ref{fig:flux_plot}.
The number quoted in Table \ref{tab:model_params} is the ML number of pulsars as determined from the likelihood distribution and the uncertainties denote the most compact 68\% confidence interval around the ML value. If the ML value for a distribution is zero, then the upper limits are given at the 68\% confidence level. In addition to the fiducial pulsar spectral index of $\alpha_{\mbox{\scriptsize psr}} = -1.7$, we include spectral indices of $-1.0$ and $-2.5$.
For comparison, the best fit parameters for the model with the number of pulsars fixed at zero is also included in Table \ref{tab:model_params} and odds ratios are calculated against this ``null'' model.
It is interesting to note that although Models A and B give only upper limits to the pulsar population (of ${\lesssim}10^3-10^4$), Model C provides a non-zero maximum likelihood value of ${\sim}10^3$. Additionally, Model C provides a better fit than the model with no pulsars at all (the ``null'' model in Table~\ref{tab:model_params}).
In all of the models considered, an increase in the maximum number of pulsars is accompanied by an increase in the free--free turnover frequency. The increased free-free absorption is required to mask the bright low-frequency tail of a large pulsar population. Current estimates of free--free absorption in the region near Sgr~A*\ give turnover frequencies around 330~MHz \citep{pae89}. Estimates of the free--free absorption of the flux from Sgr~A*, however, assume a power-law flux density for Sgr~A*\ and would likely underestimate the absorption if a pulsar population were present. As a result, any independent measurement of the free-free absorption along the line of sight of {Sgr~A*} that does not assume a spectrum of the Sgr~A*\ source could place an important constraint on the pulsar population in the inner parsecs of the Galaxy.
From current radio measurements of the inner parsec of the GC, total pulsar populations (that is, both CPs and MSPs) of up to ${\sim}10^3$ are consistent with observations, regardless of the underlying spatial distribution.
\resizebox{0.5\textwidth}{!}{
\begin{threeparttable}
\caption{Number of Pulsars within 1~pc for Given Model and Spectral Index}
\label{tab:model_params}
\centering
\begin{tabular}{cccccccc}
\toprule
\toprule
& & \multicolumn{3}{c}{Model Parameters\tmark[a]} & & &\\
\cmidrule(l){3-5}
Model & $\alpha_{\mbox{\scriptsize psr}}$ & $N_1 (\times10^3)$ & $\alpha_{\mbox{\scriptsize bh}}$ & $\nu_f$~(GHz) & $N_{dof}$ & $\chi^2_r$ & Odds\\
\cmidrule(l){1-8}
A & $-1.0$ & $<16.2$ & 0.14 & 0.37 & 10 & 1.08 & $10^{-0.64}$\\
& $-1.7$ & $<7.4$ & 0.14 & 0.38 & 10 & 1.16 & $10^{-1.05}$\\
& $-2.5$ & $<2.2$ & 0.14 & 0.38 & 10 & 1.17 & $10^{-1.35}$\\
& & & & & & & \\
B-0 & $-1.0$ & $5.3^{+2.5}_{-5.3}$ & 0.17 & 0.45 & 10 & 0.79 & $10^{-0.55}$\\
& $-1.7$ & $<4.9$ & 0.15 & 0.53 & 10 & 1.07 & $10^{-1.21}$\\
& $-2.5$ & $<0.9$ & 0.13 & 0.47 & 10 & 1.27 & $10^{-1.42}$\\
& & & & & & & \\
B-1 & $-1.0$ & $<3.8$ & 0.14 & 0.40 & 10 & 1.06 & $10^{-0.85}$\\
& $-1.7$ & $<1.5$ & 0.13 & 0.41 & 10 & 1.14 & $10^{-1.03}$\\
& $-2.5$ & $<0.4$ & 0.14 & 0.38 & 10 & 1.17 & $10^{-1.22}$\\
& & & & & & & \\
B-2 & $-1.0$ & $<2.4$ & 0.14 & 0.36 & 10 & 1.07 & $10^{-1.02}$\\
& $-1.7$ & $<1.2$ & 0.14 & 0.38 & 10 & 1.13 & $10^{-1.00}$\\
& $-2.5$ & $<0.4$ & 0.14 & 0.38 & 10 & 1.17 & $10^{-1.23}$\\
& & & & & & & \\
C & $-1.0$ & $1.4^{+0.5}_{-0.7}$ & 0.25 & 0.40 & 10 & 0.55 & $10^{+0.28}$\\
& $-1.7$ & $1.1\pm 0.4$ & 0.21 & 0.47 & 10 & 0.59 & $10^{+0.12}$\\
& $-2.5$ & $1.1^{+0.4}_{-0.6}$ & 0.17 & 0.58 & 10 & 0.75 & $10^{-0.46}$\\
& & & & & & & \\
Null\tmark[b] & $\dots$ & 0 & 0.15 & 0.26 & 11 & 0.81 & 1.0\\
\bottomrule
\end{tabular}
\begin{tablenotes}
\footnotesize
\item[a] Best fit model parameters for maximum likelihood (ML) number of pulsars. If the ML number of pulsars is zero, then the 68\% confidence upper limit is used and reported with a ``$<$''.
\item[b] The ``Null'' case fixes the number of pulsars at zero.
\end{tablenotes}
\end{threeparttable}}
\begin{figure}[h!]
\centering
\includegraphics[width=0.5\textwidth]{f4_color.eps}
\caption{\emph{Top:} Fit curves for model pulsar distributions to the VLA (squares) and GMRT (diamond) data from \citet{an05} and VLA (filled circles) data from \citet{falcke98}. Since the best fit for Models A and B-1 indicate that the most probable number of pulsars is zero, we have instead plotted the best fit assuming the 68\% confidence upper limit number of pulsars are present. The upper axis gives the projected radial distance of the effective resolution beam at the distance of Sgr~A*~using the VLA (for details, see Section~\ref{ssec:ang_res}). Note that although the fits were made using both the VLA and GMRT data, the curves above only apply to the VLA data points (squares and filled circles). \emph{Middle and Bottom:} Components to the observed flux of Sgr~A*\ from the MBH point source and surrounding pulsars for Models A and C. Note how the pulsar component in Model A rises quickly with decreasing frequency as a result of both the pulsar spectrum and the increasing number of pulsars in the beam. The total unabsorbed flux is also shown for comparison.}
\label{fig:flux_plot}
\end{figure}
\section{Diffuse Gamma-Ray Emission and MSPs}
\label{diffuse_gamma}
MSPs are known gamma-ray sources \citep{abdo09}. As a result, we may set constraints on the MSP population in the GC by measuring the diffuse gamma-ray emission from the GC. In a recent analysis of the first two years of data from the Fermi Gamma-Ray Space Telescope, \citet[herafter, HG]{hg11} observed an excess of gamma-ray flux toward the inner $1^\circ$ (150~pc) of the GC, with a significant excess within $0.25^\circ$ ($\approx40$~pc). HG argued that the signal is consistent with the annihilation of 7--10 GeV dark matter particles with a cusped halo distribution around Sgr~A* and difficult to explain using known astrophysical sources. \citet{aba11}, however, claimed that while the spectrum may be inconsistent with the \emph{average} pulsar spectrum, it is consistent with \emph{some} pulsar spectra and therefore could be explained by astrophysical sources. HG provide a spectrum for the gamma-ray excess in the GC of the form
\begin{equation}
\frac{dN_\gamma}{dE} \propto E^{-\Gamma} \exp \left( -E / E_{cut}\right)
\end{equation}
where $\Gamma = 0.99^{+0.10}_{-0.09}$ and $E_{cut} = 1.92^{+0.21}_{-0.17}$~GeV. These parameter values are consistent (to current uncertainties) with the spectra of 16 out of 46 pulsars in the first Fermi LAT catalog of gamma-ray pulsars \citep{abdoPSR10}. We proceed assuming that the observed excess seen by HG is real \citep[however, see ][]{bmr11}, follows the spectrum fit by HG and is entirely caused by a collection of MSPs in the GC.
The spectrum is normalized so that
$E^2 dN_\gamma/dE \approx 10^{-7}~\mbox{GeV}~\mbox{cm}^{-2}~\mbox{s}^{-1}$
at $E=1.0$~GeV based on the HG plots, giving
\begin{eqnarray}
\frac{dN_\gamma}{dE} &=& 1.7 \times 10^{-7} \mbox{ GeV}^{-1} \mbox{cm}^{-2} \mbox{s}^{-1} \left(\frac{E}{1 \mbox{ GeV}}\right)^{-0.99} \nonumber \\
& &\times\exp{\left( -\frac{E}{1.92 \mbox{ GeV}} \right)}.
\end{eqnarray}
Calculating the integrated energy flux from the above spectrum over 0.1--100~GeV as
\begin{eqnarray}
S_{\gamma} = \int_{0.1~\mbox{\scriptsize GeV}}^{100~\mbox{\scriptsize GeV}} E \frac{dN_\gamma}{dE} dE
\end{eqnarray}
and letting $L_{\gamma} = 4\pi d^2 f_{\Omega} S_{\gamma}$, we find that the gamma-ray luminosity of a source at the GC is $L_{\gamma} \approx 4 \times 10^{36} \mbox{ erg s}^{-1} f_{\Omega}$. The correction factor, $f_{\Omega}$, is similar to the beaming fraction in radio pulsars and is generally taken to be unity in most modern models \citep{wrw09}.
Following calculations made to estimate the number of MSPs in globular clusters \citep{abdoGC10}, we can estimate the MSP population of the GC as
\begin{equation}
\label{gamma_est}
N_{\mbox{\scriptsize MSP}} = \frac{L_{\gamma}}{\langle \dot{E} \rangle \langle \eta_{\gamma} \rangle}
\end{equation}
where $\dot{E}$ is the total spindown luminosity of a MSP, $\eta_{\gamma} = L_{\gamma}/\dot{E}$ is the ``efficiency'' of converting spin-down power into gamma-rays, and angled brackets denote an average over the population.
The values of $\langle \dot{E} \rangle$ and $\langle \eta_{\gamma} \rangle$ are taken from the local ($d<1$~kpc) MSP population to avoid selection effects.
The mean spin-down luminosity is taken to be $\langle \dot{E} \rangle = 1.1 \times 10^{34} \mbox{ erg s}^{-1}$ from the 27 MSPs within 1~kpc of Earth listed in the ATNF catalog \citep{PSRCAT}.
Since
$\langle \eta_{\gamma} \rangle \propto L_\gamma \propto d^2$,
the uncertainties in $\langle \eta_{\gamma} \rangle$ are dominated by distance uncertainties and values range from $\langle \eta_{\gamma} \rangle = 10^{-3}-1$ in the first \emph{Fermi} gamma-ray pulsar catalog \citep{abdoPSR10}.
To mitigate this distance problem, we take only those seven nearby pulsars in the catalog for which distances could be measured accurately with parallax.
For these pulsars, we find $\langle \eta_{\gamma} \rangle = 0.08\pm 0.04$.
Using Equation~\ref{gamma_est}, we get an estimate of $N_{\mbox{\scriptsize MSP}} \approx 5000$.
In order to contribute the observed excess, these pulsars would be located within $1^{\circ}$ ($150$ pc) of Sgr~A*, with the highest concentration within $0.25^{\circ}$ ($40$ pc).
Our estimate for the number of MSPs in the GC is essentially an upper bound for the population. However, modeling the background component of the diffuse gamma-ray emission in the GC is still somewhat uncertain \citep{abdoSRC09} and, as a result of this uncertainty, our estimate can only be taken as an approximate upper limit. Additionally, as this is a measure of the excess gamma-ray flux in the region (with the Galactic plane and a central point source coincident with Sgr~A*\ subtracted), there exists the possiblity that a signicant number of MSPs are unaccounted for in this estimate. Assuming the HG gamma-ray excess in the GC is real and does not suffer from systematic errors in background subtraction, we see that it is not inconsistent with a centrally concentrated population of ${\sim}10^3$ MSPs in the inner tens of parsecs from Sgr~A*.
Finally, we note that similar estimates for the MSP population in the GC have been made in the past. \citet{wjc05} used a model proposed by \citet{zc97} to predict the emission of gamma-rays from MSPs as a function of global pulsar parameters like spin period and polar magnetic field. They used Monte Carlo methods to simulate a population of pulsars and measured the total gamma-ray luminosity. Comparing this luminosity to measured values from EGRET, \citet{wjc05} estimated a GC MSP population of $N_{\mbox{\scriptsize MSP}}\approx6000$ in the EGRET field ($r \approx 1.5^{\circ}$, 220~pc).
\section{Massive Stars in the Galactic Center}
\label{massive_stars}
The central parsec of the GC is one of the most active massive star formation regions in the Milky Way and is currently known to contain about 200 young massive stars \citep{geg10}. As massive stars are the progenitors of neutron stars (NSs), we may use current stellar populations to estimate the number of pulsars in this region.
The inner parsec stellar population is divided into three fairly distinct regions. The innermost region $\left(R\leq1^{\prime\prime}\right)$ is the so-called ``S-star Cluster'' of main sequence B-stars. These stars can be fitted with a standard Salpeter initial mass function (IMF) for a single star formation event or for a continuous star-forming population with ages of a few Myr to 60 Myr \citep{bartko10}. Outside this region $\left( 1^{\prime\prime}\leq R\leq 12^{\prime\prime}\right)$, the stars are largely arranged in at least one disk (possibly two) of mass $M\sim 10^4 M_{\odot}$ with a top-heavy IMF of $dN/dm \propto m^{-0.45}$ \citep{bartko10}. The early-type stars in this region appear to have been formed in a starburst ${\sim}6$ Myr ago. Outside the disk region $\left( R \geq 12^{\prime\prime}\right)$ the stellar population is again consistent with a Salpeter IMF.
Following similar calculations by \citet{lc08} and \citet{fl11}, the current stellar populations can be used to estimate the number of radio pulsars harbored in the central parsec. The number of active CPs and MSPs beamed toward Earth are estimated to be
\begin{eqnarray}
\label{eq:cppops}
N_{\mbox{\scriptsize {CP}}} = f_{psr} f_{b} f_{\tau} f_{v} N_{ns}
\end{eqnarray}
and
\begin{eqnarray}
\label{eq:msppops}
N_{\mbox{\scriptsize MSP}} = f_{psr} f_{b} f_{\tau} f_{v} f_{r} N_{ns},
\end{eqnarray}
respectively, where $N_{ns}$ is the number of NSs in the inner parsec, $f_{psr}$ is the fraction of NSs that form pulsars, $f_{b}=0.2$ is the beaming fraction, $f_{\tau}$ is the fraction of pulsars with ages less than the typical pulsar radio lifetime, $f_{v}$ is the fraction of pulsars with birth velocities small enough to be retained in the inner parsec and $f_{r}$ is the fraction of NSs that are recycled into MSPs. The fraction of NSs that form pulsars is taken to be $f_{psr}\sim1$, although this factor is still fairly uncertain \citep{lc08}. The terms $N_{ns}$, $f_{\tau}$, $f_{v}$ and $f_{r}$ are discussed below.
\subsection{Neutron Star Population in the Inner Parsec ($N_{ns}$)}
The total number of NSs residing in the inner parsec may be estimated from current observations of the massive star population. Velocity measurements of stars allow for the total dynamical mass enclosed within a few parsecs of {Sgr~A*} to be determined. Subtracting the mass of the MBH gives an extended mass of ${\sim}10^6~M_{\odot}$ in stars within a parsec of {Sgr~A*} \citep{gtk96, sme09}. With the total mass of stars established, the number of NSs can be calculated for a given initial mass function (IMF) and mass range of stars that end their lives as NSs.
A typical NS progenitor mass range is $9M_{\odot}<M<25M_{\odot}$ \citep{hfw03}. However, such ranges are theoretically determined for only isolated non-rotating stars and rely on mass-loss and stellar wind models that are poorly constrained by observations \citep{hfw03}. If very massive stars have higher mass-loss rates than modeled, then NSs could form from stars with initial masses above $\gtrsim25M_{\odot}$. That this may be the case is supported by the limited observational constraints. For example, \citet{mcc06} detected an X-ray pulsar in the young Galactic cluster Westerlund~1 that requires a progenitor mass of $M>40M_{\odot}$ to have formed in the age of the cluster.
To calculate the number of NSs produced by the GC massive star population, we consider IMFs of the form $dN/dm\propto m^{-\alpha}$ over a range of masses from $0.1M_{\odot}$ to $100M_{\odot}$. For the case of the standard Salpeter IMF ($\alpha=2.35$) observed in most of the inner parsec and a progenitor mass range of $9M_{\odot}<M<25M_{\odot}$, a total of $N_{ns}\approx4900$ NSs are produced. If the top-heavy IMF ($\alpha=0.45$) observed in the young disks is taken to hold for the whole central parsec, a total of $N_{ns}\approx5700$ are produced. If a wider range of progenitor masses is taken, say $9M_{\odot}<M<40M_{\odot}$ to allow the \citet{mcc06} observation, the total NS populations increase to $N_{ns}\approx5700$ and $N_{ns}\approx9500$ for the Salpeter and top-heavy IMFs, respectively.
Even given a wide range in IMF and progenitor mass, the above results are consistent to an order of magnitude with a NS population of $N_{ns}\sim10^4$. In terms of NSs formed per stellar mass, we find a value of $\beta_{ns}\sim10^{-2}M^{-1}_{\odot}$ for the ${\sim}10^6M_{\odot}$ worth of stars within a parsec of Sgr~A*.
\subsection{Fraction of Still-Active Pulsars ($f_{\tau}$)}
\label{ss:f_r}
The typical lifetimes over which pulsars maintain active radio emission are $\tau\sim10^7$~yr for CPs and $\tau\sim10^9-10^{10}$~yr for MSPs. As a result, the fraction of pulsars formed recently enough to still be active must be considered in estimates of the observable population. This factor will depend on the star formation history (SFH) of the region. In the case of continuous star formation, the active fraction of pulsars can be estimated to be $f_{\tau}\sim\tau/t_{sf}$, where $t_{sf}$ is the amount of time elapsed since the star formation began.
\subsection{Fraction of Pulsars Retained ($f_{v}$)}
\label{ss:fv}
From the high observed velocities (${\sim}10^2-10^3~\mbox{km~s}^{-1}$) of some pulsars, it has been inferred that NSs are given a large ``kick'' velocity at birth as a result of binary disassociation or an asymmetric supernova explosion (or both). A pulsar created in the inner parsec will be retained in the inner parsec only if its birth velocity does not exceed the local escape velocity of its orbit around Sgr~A*. Assuming only the influence of a MBH of mass $M\approx 4 \times 10^6~\mbox{M}_{\odot}$, the escape velocity at a distance $r$ from Sgr~A* is given by
\begin{equation}
\label{escape_vel}
v_e(r) = 185~\mbox{km s}^{-1}~\left(\frac{M}{4\times10^6~\mbox{M}_{\odot}}\right)^{1/2}\left(\frac{r}{1~\mbox{pc}}\right)^{-1/2}.
\end{equation}
Assuming a Maxwellian distribution of birth velocities with a mean of $\langle v_{birth} \rangle = 380~\mbox{km s}^{-1}$ \citep{fgk06}, we find retention fractions of $f_v \approx 0.1$ and $f_v \approx 0.25$ for radial distances of $r=1~\mbox{pc}$ and $r=0.5~\mbox{pc}$, respectively. However, the actual shape of the pulsar birth velocity distribution is not well constrained and one may worry that the Maxwellian distribution is arbitrary. In an analysis of pulsar velocities, \citet{fgk06} consider six different pulsar velocity distributions. Repeating our calculation for each distribution, we find retention fractions in the ranges of $f_v \approx 0.05 - 0.4$ and $f_v \approx 0.1 - 0.5$ for distances of $r = 1~\mbox{pc}$ and $r=0.5~\mbox{pc}$, respectively.
If no other effects are important, we would expect a retention fraction of $f_v\gtrsim0.1$. However, a similar analysis to the one performed above would underestimate the retained NS populations in globular clusters by orders of magnitude. Observations of pulsars and low-mass X-ray binaries (LMXBs) suggest that up to ${\sim}10\%$ of all NSs formed in some globular clusters may be retained \citep{prp02}. However, current models for isolated pulsars predict that $\lesssim1\%$ of NSs will have birth velocities below the $\lesssim50~\mbox{km s}^{-1}$ globular cluster escape velocities. This ``retention problem'' is currently an unsolved problem, but likely has to do with the high stellar densities and binary fractions found in the cores of globular clusters \citep[see][and references within]{prp02}. Since the GC has even higher stellar densities than cores of globular clusters, we expect a similar heightening of the retention fraction and thus adopt as a nominal value $f_v \sim 1$.
\subsection{Fraction of Pulsars Recycled to MSPs ($f_{r}$)}
\label{ssec:f_r}
MSPs are thought to be formed when a NS in a binary gains angular momentum through accretion of matter in a process known as ``recycling'' \citep{acr82,bv91}. Thus, the fraction of NSs recycled into MSPs must be determined to estimate the MSP population. Within 3~kpc of the Sun, the birthrate of pulsars with 400~MHz pseudo-luminosities above $1~\mbox{mJy}~\mbox{kpc}^2$ is observed to be ${\sim}10^{-3}~\mbox{yr}^{-1}$ and ${\gtrsim}10^{-6}~\mbox{yr}^{-1}$ for CPs and MSPs, respectively \citep{parkesII1997}. Using these birthrates, we may infer that the recycling fraction is at least $f_r\gtrsim10^{-3}$ in the Galactic disk. The increased stellar density and stellar encounter rate in the GC will very likely increase this fraction. LMXBs, the assumed progenitors of MSPs, have been found to be ${\sim}100$ times more abundant in globular clusters than the general Galactic field \citep{clark75, katz75}. As the central parsec of the GC has a higher stellar density than globular clusters, we would expect at least a similar overabundance of LMXBs and their resultant MSPs as is seen in globular clusters. As a result, we adopt a recycling fraction of $f_r\sim0.1$.
\subsection{Pulsar Estimates for Various Star Formation Histories}
As the star formation history (SFH) of the central parsec of the GC is still somewhat uncertain, we will consider two general SFHs suggested by current observations. In the first case, we take the massive-star disk(s) to have formed in a well-defined starburst ${\sim}6$~Myr ago \citep{pgm06,bartko10} and assume that the rest of the central parsec has experienced continuous star formation over the age of the Galaxy. In the second case, we consider SFHs based on spectro-photometry of cool giant stars which indicate that most of the stars in the inner parsec were formed $\gtrsim5$~Gyr ago but also show an increased star formation rate in the last ${\sim}100$~Myr \citep{brs03, pfz11}.
\subsubsection{Continuous Star Formation + Disk Starburst}
In the first SFH, pulsars can come from both the general population of stars in the central parsec and the young disk population. For the general population, we take the total mass of stars to be $M\sim10^6{M}_\odot$ and assume continuous star formation over the last ${\sim}10^{10}$~yr. From the parameters discussed above, the total number of NSs is found to be
$N_{ns}\sim10^4\left(\beta_{ns}/10^{-2}\mbox{M}^{-1}_\odot\right)$.
Taking a CP active radio lifetime of $\tau\sim10^7$~yr, continuous star formation over ${\sim}10^{10}$~yr will give the fraction of CPs still active to be $f_{\tau}\sim10^{-3}$. From Equation~\ref{eq:cppops}, the CP contribution from the continuous star forming region of the inner parsec is
\begin{eqnarray}
N^{\mbox{\scriptsize {gen}}}_{\mbox{\scriptsize {CP}}}\sim2\left(\frac{f_v}{1.0}\right)\left(\frac{\beta_{ns}}{0.01~\mbox{M}^{-1}_\odot}\right)\left(\frac{M}{10^6~\mbox{M}_\odot}\right).
\end{eqnarray}
For MSPs with a radio lifetime of ${\sim}10^{10}$~yr, the fraction still active is $f_{\tau}\sim1$. Adopting the recycling fraction of $f_r\sim0.1$ discussed above, the contribution of MSPs from the general population of the inner parsec is given by Equation~\ref{eq:msppops} as
\begin{eqnarray}
N^{\mbox{\scriptsize {gen}}}_{\mbox{\scriptsize {MSP}}}{\sim}200\left(\frac{f_v}{1.0}\right)\left(\frac{f_{r}}{0.1}\right)\left(\frac{\beta_{ns}}{0.01~\mbox{M}^{-1}_\odot}\right)\left(\frac{M}{10^6~\mbox{M}_\odot}\right).
\end{eqnarray}
Additionally, the contribution of pulsars from the massive-star disk(s) must be considered. The disk is assumed to have a population of stars formed in a starburst event ${\sim}6\times10^6$~yr ago with a total stellar mass of ${\sim}10^4{M}_{\odot}$ \citep{pgm06,bartko10}. Since the age of the disk is comparable to the active radio lifetime of a CP, the fraction of CPs still active is taken to be $f_{\tau}\sim1$. From Equation~\ref{eq:cppops}, the CP contribution from the disk is found to be
\begin{eqnarray}
N^{\mbox{\scriptsize {disk}}}_{\mbox{\scriptsize {CP}}}\sim20\left(\frac{f_v}{1.0}\right)\left(\frac{\beta_{ns}}{0.01~\mbox{M}^{-1}_\odot}\right)\left(\frac{M_{disk}}{10^4~\mbox{M}_\odot}\right).
\end{eqnarray}
The disk population is not expected to produce any currently observable MSPs as the short timescale of ${\sim}6\times10^6$~yr provides insufficient time to create and evolve a NS population into MSPs. As a result, this first SFH produces roughly $N_{\mbox{\scriptsize {CP}}}\sim20$ CPs and $N_{\mbox{\scriptsize MSP}}\sim200$ MSPs.
\subsubsection{SFH from Observations of Cool Giant Stars}
The SFH has also been estimated by comparing simulated populations with the observed cool giant stars in the central parsec. Such simulations allow the average star formation rate to be calculated as a function of look-back time for a few coarse time bins. In two separate analyses, both \citet{brs03} and \citet{pfz11} found that ${\gtrsim}80\%$ of the stellar mass in the central parsec was formed ${\gtrsim}5$~Gyr ago and that there has been increased star formation in the last ${\sim}100$~Myr. In their best fit models, \citet{brs03} found an average star formation rate of ${\sim}3\times10^{-3}M_{\odot}\mbox{yr}^{-1}$ within 2~pc of {Sgr~A*} from 10 to 100~Myr ago and \citet{pfz11} found an average star formation rate of ${\sim}10^{-3}M_{\odot}\mbox{yr}^{-1}$ within 1~pc of {Sgr~A*} from 50 to 200~Myr ago. Both cases are consistent with ${\sim}10^5M_{\odot}$ worth of stars being formed in the inner parsec of the GC in the last ${\sim}100$~Myr.
If there has been continuous star formation in the last ${\sim}100$~Myr, then the fraction of still active CPs would be $f_{\tau}\sim0.1$. Taking all other parameters as before, the continuous formation of ${\sim}10^5M_{\odot}$ worth of stars over the last ${\sim}10^8$~yr would produce
\begin{eqnarray}
N^{\mbox{\scriptsize {con}}}_{\mbox{\scriptsize {CP}}}\sim20\left(\frac{f_v}{1.0}\right)\left(\frac{\beta_{ns}}{0.01~\mbox{M}^{-1}_\odot}\right)\left(\frac{M\left(t<10^8~\mbox{yr}\right)}{10^5~\mbox{M}_\odot}\right).
\end{eqnarray}
If the recent star formation all took place in the last ${\sim}10^7$~yr, then the fraction of CPs still active would be $f_{\tau}\sim1$. In this case, the number of active CPs would be
\begin{eqnarray}
N^{\mbox{\scriptsize {burst}}}_{\mbox{\scriptsize {CP}}}{\sim}200\left(\frac{f_v}{1.0}\right)\left(\frac{\beta_{ns}}{0.01~\mbox{M}^{-1}_\odot}\right)\left(\frac{M\left(t<10^8~\mbox{yr}\right)}{10^5~\mbox{M}_\odot}\right).
\end{eqnarray}
In either of the above cases, the majority $\left({\gtrsim}90\%\right)$ of the star formation took place at look-back times ${\gtrsim}10^8$~yr ago. As a result, the number of MSPs produced will be approximately the same as the first SFH considered, namely $N_{\mbox{\scriptsize MSP}}\sim200$.
\subsubsection{Upper Limits to the Pulsar Population}
Using the above estimates of CP and MSP populations for a range of observationally supported SFHs, upper limits may be set on the total allowable number of pulsars in the inner parsec. For CPs, the most favorable formation scenarios produce $N_{\mbox{\scriptsize {CP}}}\sim200$. For MSPs, a variety of SFHs consistently produce $N_{\mbox{\scriptsize MSP}}\sim200\left(f_r/0.1\right)$. Since the recycling fraction is unknown for the extreme conditions of the inner GC, an upper limit of $N_{\mbox{\scriptsize MSP}}\sim2000$ may be set by adopting $f_r\sim1$. Thus, observations of current stellar populations place an upper limit of $\mbox{a few}\times10^3$ on the number of active radio pulsars beamed toward Earth in the inner parsec of the GC.
\section{Pulsar Wind Nebulae in Inner 20 pc}
\label{PWN_sec}
Using a total of 1~Ms of \emph{Chandra} ACIS--I observations of the inner parsecs of the GC, \citet{mbb08} compiled a catalog of 34 diffuse X-ray emitting features. Based on the X-ray luminosities and sizes of the sources in their catalog, \citet{mbb08} expect ${\sim}20$ pulsar wind nebulae (PWNe) to be present within 20~pc of Sgr~A*. Since PWNe are powered by pulsars, we may use this inferred population of PWNe to estimate the pulsar population in the inner 20~pc of the GC.
Pulsars can lose their rotational kinetic energy by the release of relativistic winds of charged particles. The winds exert a pressure upon and deposit energy into the surrounding interstellar medium, producing luminous PWNe that radiate across the electromagnetic spectrum \citep[see, e.g.,][]{gs06}. As a result, the luminosity of the PWN will be directly related to the spin-down luminosity of the pulsar given by
\begin{equation}
\label{edot}
\dot{E} = -dE_{rot}/dt = 4\pi^2I\dot{P}/P^3,
\end{equation}
where \emph{I} and \emph{P} are the moment of inertia and period, respectively, of the pulsar.
Since the spin-down luminosity of a pulsar decreases with increasing age, one would expect the most luminous PWNe to contain young pulsars \citep[however, older pulsars may be ``recycled'' sufficiently to power PWNe as described by][]{ctw06}. Of the 30 confirmed pulsars associated with PWNe, 25 have characteristic ages ($\tau_c = P/2\dot{P}$) of $\tau_c \lesssim 10^6~\mbox{yr}$ and 23 have characteristic ages $\tau_c \lesssim 10^5~\mbox{yr}$ \citep{roberts04}. Thus, we adopt a lifetime for a typical PWN of $t_{pwn} \sim 10^5~\mbox{yr}$.
Given the observed number of PWNe in the GC ($N_{obs} \sim20$) and a typical lifetime of $t_{pwn} \sim 10^5~\mbox{yr}$, we find a mean rate of formation of PWNe over the last ${\sim}10^5~\mbox{yr}$ to be
\begin{equation}
\label{pwn_rate}
\beta_{pwn} \sim 2\times10^{-4}~\mbox{yr}^{-1}\left(\frac{N_{pwn}}{20}\right)\left(\frac{t_{pwn}}{10^5~\mbox{yr}}\right)^{-1}.
\end{equation}
Assuming that the PWN formation rate has remained constant over the last ${\sim}10^7~\mbox{yr}$, we may estimate the number of CPs in the inner 20~pc to be
\begin{equation}
\label{PWN_cp}
N_{\mbox{\scriptsize {CP}}} \sim \beta_{pwn} \tau_{psr} f_b f_v f_{pwn}^{-1},
\end{equation}
where $\tau_{psr}\sim10^7~\mbox{yr}$ is the typical radio lifetime of a CP, $f_b=0.2$ is the beaming fraction, $f_v$ is the fraction of pulsars with birth velocities low enough to be retained in the inner 20~pc, and $f_{pwn}\sim1$ is the fraction of pulsars that form PWNe.
Taking the mass of the central 20~pc to be $M=3\times10^7~M_{\odot}$ \citep{lhw92}, we find that pulsars must have velocities $v_{birth} < 115~\mbox{km~s}^{-1}$ to remain gravitationally bound to the inner 20~pc. From the birth velocity distributions considered by \citet{fgk06}, we find $f_v\approx0.05-0.30$. However, the distance traveled by a pulsar is given by
\begin{equation}
\label{dist_psr}
d \approx 10~\mbox{pc}\left(\frac{v}{100~\mbox{km s}^{-1}}\right)\left(\frac{t}{10^5~\mbox{yr}}\right).
\end{equation}
Thus, pulsars with velocities high enough to become gravitationally unbound will also have velocities high enough to escape the inner 20~pc on timescales comparable to the PWN lifetime. As a result, the PWNe observed within the inner 20~pc are very likely to remain there and we can take $f_v\sim1$.
From Equation~\ref{PWN_cp}, we see that if the PWN birth rate has been constant over the last ${\sim}10^7~\mbox{yr}$, we would expect
\begin{equation}
\label{npsr_pwn}
N_{\mbox{\scriptsize {CP}}} \sim 400 \left(\frac{N_{pwn}}{20}\right)\left(\frac{t_{pwn}}{10^5~\mbox{yr}}\right)^{-1}\left(\frac{f_{pwn}}{1.0}\right)^{-1}
\end{equation}
CPs within 20~pc of Sgr~A*.
Finally, we note that of the ${\sim}20$ PWN candidates identified by \citet{mbb08}, 4 fall within a projected radial distance of 1~pc from Sgr~A*. Assuming the number of pulsars scales accordingly, then Equation~\ref{npsr_pwn} predicts $N_{CP}\sim80$ CPs within the inner parsec of the GC.
\section{Supernova Rate in the Galactic Center}
\label{supernova}
Neutron stars are formed as the end products of core-collapse supernovae (CCSN). An estimate of the rate of CCSN in the GC would therefore offer a constraint on the pulsar population. The CCSN rate is estimated below for both $r<150~\mbox{pc}$ and $r<20~\mbox{pc}$.
\subsection{CCSN Rate Within $r<150~\mbox{pc}$ of Sgr~A*}
By measuring the total mass of ${}^{26}\mbox{Al}$ in the Galaxy, \citet{diehl06} estimate the Galactic CCSN rate to be ${\beta}_{CCSN} = 1.9 \pm 1.1 {\mbox{ century}}^{-1}$.
One may, in principle, scale this estimate to smaller regions of the Galaxy using massive star populations. Taking the inner 500 pc to contain 10\% of the Galaxy's massive star formation \citep{figer08} we can estimate that the inner ${\sim}150$ pc contains ${\sim}2\%$ of the massive star formation and therefore should have a CCSN rate of
${\beta}_{CCSN} \approx 0.04 {\mbox{ century}}^{-1}$. \citet{crocker11} estimate a similar rate and show that it is consistent with SN rate estimates from infrared observations, stellar composition, X-ray emission, gas turbulence, and high-velocity compact clouds \citep[see][and references within]{crocker11}.
We may now estimate the CP population in the GC to be
\begin{eqnarray}
N_{\mbox{\scriptsize {CP}}} = f_{psr}f_{b}f_{v}{\tau}_{psr}{\beta}_{CCSN}
\end{eqnarray}
where
${\beta}_{CCSN} \approx 4 \times 10^{-4} {\mbox{yr}}^{-1}$ is the CCSN rate,
${\tau}_{psr} \sim 10^7 \mbox{ yr}$ is the mean canonical pulsar lifetime,
$f_b=0.2$ is the fraction of pulsars beamed toward Earth,
$f_{v}$ is the fraction of pulsars with birth velocities small enough to be retained by the GC and $f_{psr}\sim1$ is the fraction of CCSN that result in active pulsars. Using the distributions from \citet{fgk06}, the fraction of pulsars with birth velocities smaller than the escape velocity $v_e\approx200~\mbox{km s}^{-1}$ at 150~pc ranges from $f_v \approx 0.1-0.4$.
These values give an estimate of $N_{\mbox{\scriptsize {CP}}} \sim100$. Likewise, accounting for the longer ages for MSPs ($\tau_{psr}\sim10^{10}$~yr), we can estimate the MSP population to be $N_{\mbox{\scriptsize MSP}}\sim10^5f_r$, where $f_r$ is the fraction of NSs that get recycled to MSPs (see Section~\ref{ssec:f_r}).
\subsection{CCSN Rate in Inner 20~pc from X-ray Observations}
Studies of diffuse X-ray emission can also provide insight into the SN rate in the GC. Using over 600~ks of \emph{Chandra} ACIS--I observations, \citet{mbb04} found that the diffuse X-ray emissions in the GC could be explained by a two-temperature plasma composed of a ``soft'' component ($kT\approx0.8~\mbox{keV}$) and a ``hard'' component ($kT\approx8~\mbox{keV}$). Assuming the soft component of the plasma is primarily heated by SNe, an estimate for the SN rate can be made by observing the loss of energy from the inner 20~pc.
Let us first consider the case in which the soft component of the plasma just cools radiatively. The X-ray luminosity of the soft component of the plasma in the inner 20~pc is $L_X\approx3\times10^{36}~\mbox{erg s}^{-1}$ \citep{mbb04}. If each SN transfers ${\sim}1\%$ of its total kinetic energy of ${\sim}10^{51}~\mbox{erg}$ to the plasma, then a SN rate of $\beta_{SN}\approx10^{-5}~\mbox{yr}^{-1}$ is required to maintain the currently observed temperature. Taking this rate to be constant and $f_v\sim0.1$ (see Section~\ref{PWN_sec}) gives an estimate for the CP population of $N_{\mbox{\scriptsize {CP}}}\sim2\left(f_v/0.1\right)$.
If the plasma is unconfined, it can also cool through adiabatic expansion. Rough estimates put this cooling rate at $L_{ad}\approx9\times10^{38}~\mbox{erg s}^{-1}$ \citep{mbb04}, which would require a SN rate of $\beta_{SN}\approx3\times10^{-3}~\mbox{yr}^{-1}$. Again assuming this rate is constant and $f_v\sim0.1$ as above, the estimated CP population is $N_{\mbox{\scriptsize {CP}}}\sim600\left(f_v/0.1\right)$.
The above estimates assume that the SN rate is constant over the radio lifetime of a CP (${\sim}10^7$~yr). However, the SN rate estimates are only required to hold over a characteristic cooling time of $t_{c}\sim E/L$, where $E\sim5\times10^{50}~\mbox{erg s}^{-1}$ is the total thermal energy stored in the plasma and $L$ is the appropriate cooling luminosity \citep{mbb04}. The cooling timescales for radiative cooling and adiabatic expansion are $5\times10^6$~yr and $2\times10^4$~yr, respectively.
Keeping this caveat in mind, the assumption of a constant SN rate does allow useful bounds to be put on the pulsar populations in the GC. Taking a constant SN rate from radiative cooling to be a lower bound and a constant SN rate from adiabatic expansion to be an upper bound, we find the CP population in the inner 20~pc to be $1\lesssim N_{\mbox{\scriptsize {CP}}}\lesssim10^3$. Likewise, for MSPs with $\tau\sim10^{10}$~yr, we estimate a population of $10^3f_r\lesssim N_{\mbox{\scriptsize MSP}}\lesssim10^6f_r$, where $f_r$ is the fraction of NSs recycled into MSPs (see Section~\ref{ssec:f_r}).
\section{Discussion}
\label{discussion}
\subsection{Summary of Estimates}
We have used observations over a wide range of wavelengths to make order of magnitude estimates of the number of pulsars allowed within $r\leq1$~pc and $r\leq150$~pc of Sgr~A*. The estimates are summarized in Table~\ref{summary1}.
\begin{table*}[h!t]
\begin{center}
\begin{threeparttable}
\caption{Summary of Pulsar Population Estimates by Method}
\label{summary1}
\begin{tabular}{lcccccc}
\toprule
\toprule
& \multicolumn{3}{c}{$r\leq1~\mbox{pc}$} & \multicolumn{3}{c}{$r\leq150~\mbox{pc}$}\\
\cmidrule(l){2-4} \cmidrule(l){5-7}
Method & CP & MSP & Total & CP & MSP & Total\\
\midrule
Pulsar Surveys & ${\lesssim}10^3$ & $\dots$ & $\dots$ & $\dots$ & $\dots$ & ${\gtrsim}10^2$\\
Radio Point Sources & $\dots$ & $\dots$ & $\dots$ & $\dots$ & $\dots$ & ${\lesssim}10^4$\\
Radio Spectrum A& $\dots$ & $\dots$ & $< 7.4\times10^3$ & $\dots$ & $\dots$ & $\dots$\\
Radio Spectrum B-1& $\dots$ & $\dots$ &$<1.5\times10^3$ & $\dots$ & $\dots$ & $\dots$\\
Radio Spectrum C& $\dots$ & $\dots$ &$(1.1\pm0.4)\times10^3$ & $\dots$ & $\dots$ & $\dots$ \\
Diffuse Gamma-ray & $\dots$ & $\dots$ & $\dots$ & $\dots$ &${\sim}5\times10^3 $& $\dots$ \\
Massive Stars\tmark[a] &${\lesssim}10^2$ & ${\lesssim}10^3$ & ${\lesssim}10^3$ & $\dots$ & $\dots$ & $\dots$\\
PWNe & $10^2$ & $\dots$ & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
Supernovae\tmark[a,b]&${\lesssim}10^3$ & ${\lesssim}10^4(f_{r}/10^{-2})$ & ${\lesssim}10^4$& $10^2$ & $10^3\left(f_{r}/10^{-2}\right)$ & $\dots$\\
\bottomrule
\end{tabular}
\begin{tablenotes}
\footnotesize
\item[a] Typical values for the ``recycling fraction'' are $f_r\gtrsim10^{-3}$ in the Galactic field and potentially as high as $f_r\gtrsim0.1$ in the central parsec (see Section~\ref{ssec:f_r})
\item[b] Supernovae estimate given at $r\leq20$~pc and $r\leq150$~pc
\end{tablenotes}
\end{threeparttable}
\end{center}
\end{table*}
\subsubsection{Pulsar Population Within $r\leq150$~pc}
Limits on the pulsar population within 150~pc of Sgr~A*\ may be set using the five known pulsars in the inner $15\arcmin$, the catalog of compact radio sources by \citet{lc08}, measurements of an excess gamma-ray flux by \citet{hg11} and estimates of the Galactic core-collapse supernova rate.
The five known pulsars in the inner $15\arcmin$ provide the strongest current evidence for an intrinsic pulsar population in the GC region. A Monte Carlo population analysis by \citet{dcl09} showed that the pulsar detections in the survey by \citet{dcl09} indicate a population of at least $N\gtrsim100$ pulsars within 100~pc of Sgr~A*.
The catalog of compact radio sources compiled by \citet{lc08} allows for upper limits to be placed on the pulsar population in the inner 150~pc. The VLA survey produced a total of 170 compact steep-spectrum sources. Although pulsars cannot be unequivocally classified in an imaging survey, \citet{lc08} estimate that $N\sim10$ of the sources were likely pulsars. Using this estimate and the survey sensitivities, a conservative upper bound of $N\lesssim10^4$ may be set on the pulsar population in this region. Even if it turns out that there are no pulsars in the catalog, an upper limit of $N\lesssim10^3$ may be set to 99\% confidence level. We note that although these upper limits should hold for the entire survey region, the increasing background temperature and sidelobes caused by extended emission toward the inner GC mean that the inner field of the survey (half-power radius of 15\arcmin) likely experienced decreased sensitivity and could potentially hide a significant pulsar population in the immediate vicinity of Sgr~A*.
Using the first two years of \emph{Fermi} data, \citet{hg11} claim to have detected an excess diffuse gamma-ray flux in the inner 150~pc of the Galaxy, which they attribute to annihilating dark matter particles. If we assume the gamma-rays come instead from a collection of MSPs, an upper limit to the number of MSPs in the GC may be set. Following similar calculations for globular clusters by \citet{abdoGC10}, we find that the excess is consistent with a population of ${\sim}5\times10^3$ MSPs. This estimate is nominally an upper limit, but considering the systematic uncertainties in current GC gamma-ray background models and difficulties subtracting the point source associated with Sgr~A*, we adopt this value as only a lower limit on the upper bound of MSPs in the inner 150~pc.
As radio pulsars are formed in core-collapse SNe, the SN rate also provides a constraint on the pulsar population in the GC. By scaling down the Galactic SN rate based on massive star populations in a manner similar to that of \citet{crocker11}, we estimate a SN of $\beta_{CCSN}\approx0.04~\mbox{century}^{-1}$. Such a SN rate
indicates a CP population of $N_{\mbox{\scriptsize {CP}}} \sim100$ and an MSP population of $N_{\mbox{\scriptsize MSP}} \sim10^5f_r$, where $f_r$ is the recycling fraction discussed in Section~\ref{ss:f_r}. Since the recycling fraction of field pulsars is $f_r\gtrsim10^{-3}$ and potentially as high as $f_r\sim0.1$ in globular clusters, we adopt a nominal value of $f_r\sim10^{-2}$, so $N_{\mbox{\scriptsize MSP}} \sim10^3 (f_r/10^{-2})$.
Finally, we note that all of the upper limits considered for pulsar populations on the large scale of 150~pc may not include contributions from the few inner parsecs as a result of decreased sensitivity or failure to incorporate recent starburst activity. As a result, we consider these limits mainly applicable in the region $1~\mbox{pc}\lesssim r \lesssim 150~\mbox{pc}$. The above estimates are consistent with a total pulsar population (that is, both CPs and MSPs) of $10^2\lesssim N_{psr} \lesssim 10^4$, a CP population of $N_{\mbox{\scriptsize {CP}}} \lesssim 10^2$ and an MSP population of $N_{\mbox{\scriptsize MSP}} \lesssim 10^4$.
\subsubsection{Pulsar Population Within $r\leq1$~pc}
Limits on the pulsar population in the inner parsec of the GC have been set using the non-detections of high frequency directed pulsar searches, the spectrum of Sgr~A*\ on arcsecond scales, the population and star-formation history of the massive star progenitors of NSs, the observations of ${\sim}20$ PWN candidates in the inner 20~pc, and the limits on the SN rate based on X-ray observations in the inner 20~pc.
Directed pulsar searches of the inner parsec of the Galaxy have been conducted with the GBT at frequencies of 5, 9 and 15 GHz \citep{dcl09, mkfr10}. Since none of these searches made any detections, an upper limit may be set on the total pulsar population that is still consistent with a null result. Using the survey parameters provided in the \citet{dcl09} and \citet{mkfr10} surveys, we find that up to $N_{psr}\lesssim10^3$ pulsars (both CPs and MSPs) may be present in the inner parsec.
Since the compact radio source Sgr~A*\ is broadened by interstellar scattering (${\approx}1\arcsec$ at 1~GHz), the observed flux density may actually be a combination of the emission near the MBH and a diffuse component from a population of unresolved pulsars. By requiring that this two-component system reproduce the observed spectrum of Sgr~A*, constraints may be placed on the pulsar population on arcsecond scales. We consider a variety of spatial distributions and find that a total population of ${\sim}10^3$ pulsars is consistent with flux density measurements, regardless of spatial distribution. The existence of such a large pulsar population would distort the low-frequency measurements of the intrinsic spectrum of Sgr~A*\ and the free-free absorption along the line of sight of Sgr~A*.
Upper limits to the total pulsar population can also be set by studying the populations of massive stars that end their lives as NSs. Infrared observations of the present day population of massive stars allows for estimates of the star formation history. We consider the case of two general SFHs and find that the CP population can get as high as $N_{\mbox{\scriptsize {CP}}}\lesssim200$ only under the most favorable conditions. More typical estimates for the CP population are $N_{\mbox{\scriptsize {CP}}} \sim 20$, with most of these being formed in the young disk of massive stars located ${\approx}0.5$~pc from Sgr~A*. The MSPs are less sensitive to the exact SFH and produce populations of $N_{\mbox{\scriptsize MSP}}\sim200$ with an upper limit of $N_{\mbox{\scriptsize MSP}}\lesssim2000$ for a range of reasonable SFHs.
Upper limits on the number of CPs in the inner parsecs of the GC may be set using the detection of ${\sim}20$ PWNe within 20~pc of Sgr~A*. Using the ${\sim}20$ PWN candidates compiled by \citet{mbb08} in a catalog of diffuse X-ray sources and assuming PWNe are produced at a constant rate, we find that as many as 400 CPs may reside in the inner 20~pc. If the CP distribution follows that of the PWN candidates, then as many as 80 CPs could reside in the inner parsec.
Finally, we consider measurements of the soft ($kT\approx0.8$~keV) component of the diffuse X-ray plasma in the inner 20~pc. Assuming that the plasma was heated by the transfer of kinetic energy from SNe, we may set an upper limit on the number of pulsars in this region \citep{mbb04}. By considering two different cooling regimes, we find that the most extreme cooling scenario will produce a population of up to ${\sim}10^3$ CPs and ${\sim}10^6f_r$ MSPs. Thus, we can set upper bounds of $N_{\mbox{\scriptsize {CP}}}\lesssim10^3$ and $N_{\mbox{\scriptsize MSP}} \lesssim10^4(f_r/10^{-2})$ for CPs and MSPs in the inner 20~pc.
Overall, we find that the above estimates are consistent with a CP population of $N_{\mbox{\scriptsize {CP}}} \lesssim100$ and an MSP population of $N_{\mbox{\scriptsize MSP}} \lesssim 10^3$ in the inner parsec of the GC.
\subsection{Conclusions}
Current observations of the GC are consistent with a population of up to ${\sim}10^3$ active pulsars beamed toward Earth within the central parsec around Sgr~A*.
This total population may consist of up to $N_{\mbox{\scriptsize {CP}}} \lesssim 100$ CPs and as many as $N_{\mbox{\scriptsize MSP}} \lesssim 10^3$ MSPs.
Such a population could distort the low-frequency measurements of both the spectrum of Sgr~A*\ and free-free absorption along the line of sight of Sgr~A*.
However, even with a potentially sizeable collection of pulsars, the difficult observing conditions of the inner regions of the GC will make individual detections a challenge.
The strong interstellar scattering and large pulse broadening times mean that typical pulsar periodicity searches at radio wavelengths will be sensitive to pulsars only at high observing frequencies and large bandwidths \citep{cl97}.
Even if most of the pulsars lie below the detection threshold, ``giant'' pulses intrinsic to the pulsar or as the result of an enhancement through multipath scattering may make a pulsar visible a search for single pulses. Since there may be many more MSPs than CPs, one may search for pulsars using interferometer imaging surveys of compact radio objects similar to that of \citet{lc08}.
Finally, search methods should be considered at wavelengths less affected by the scattering effects of the ISM.
Since many MSPs produce gamma-ray emission, a directed search of the GC region with the \emph{Fermi} LAT could potentially detect a pulsar in a blind periodicity search.
Despite the difficulties in finding pulsars in the GC, the detection of a pulsar orbiting Sgr~A*\ with an orbital period of $P_{orb}\lesssim100$~yr would provide an unparalleled test of gravity in the strong-field regime and could potentially allow the measurement of the spin and quadrupole moment of the MBH \citep{pl04, lw97, wk99, lwk12}. Additionally, the detection of even one pulsar in the inner few parsecs would provide an excellent probe of the magneto-ionic material, the gravitational potential, and the star formation history in the vicinity of Sgr~A*. In light of the significant scientific rewards and a potentially sizeable target population, we strongly recommend continued pulsar searches in the GC across a wide range of wavelengths.
\acknowledgments
We thank an anonymous referee for helpful comments that improved the clarity of this paper. This research was supported at Cornell University by NSF grants AST-1008213 and AST-1109411. Part of this research was conducted at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics \& Space Administration. We have also made use of NASA's Astrophysics Data System.
|
\section{Introduction}
The density structure of the Sun's atmosphere in the vicinity of the transition region is not well known, at least for flares.
Most of our knowledge is derived from empirical models \citep[see e.g.][]{VAL1981,Fontenla1993,Gabriel1976,Ewell1993}.
More recently, \citet{Aschwanden2002,Liu2006,Kontar2008a,Prato2009}, have attempted to derive the chromospheric density structure with the use of hard X-ray (HXR) emission from flare footpoints at different energies, assuming the thick-target beam model \citep{Brown1971}.
To fit their data, \citet{Aschwanden2002} have assumed that the density has a power-law shape with altitude above photosphere, whereas \citet{Kontar2008a}, working at slightly higher energies (and hence, deeper in the chromosphere), have assumed an exponential shape.
Uniform target ionization \citep{Brown1973,Kontar2002} was assumed in both case (fully-ionized for the former, fully neutral for the latter).
The Caltech Irreference Chromospheric Model \citep[CICM,][]{Ewell1993} supports a two-exponential atmosphere, with the lower component's scale height closely corresponding to the one derived by \citet{Kontar2008a},
and to the \citet[][thereafter VAL]{VAL1981} and \citet[][thereafter FAL]{Fontenla1993} models (i.e. $\sim$130 km).
These previous studies used single events.
\citet{Matsushita1992} and \citet{Sato2006} have statistically derived the altitude difference between spatially-averaged HXR emissions at different energies using {\it Yohkoh} HXT's \citep{Kosugi1992} four channels, covering an energy range between 14 to 93 keV.
We used the same technique of determining spatially-averaged centroids (center of mass) at different energies, and in a similar energy range, but with flares observed by the Ramaty High Energy Solar Spectroscopic Imager \citep[RHESSI,][]{Lin2002}, which has a much higher spectral resolution ($\sim$1 keV).
This will allow us to carry the data analysis one step further, deducing densities from the regions where non-thermal HXR emission is observed, assuming a thick-target beam model for bremsstrahlung emission.
We will first derive densities using a simple, direct method, and then try to fit a double-exponential with unit step ionization change density model.
\section{Observations} \label{sect:obs}
\subsection{Data selection}
We have selected all flares in the RHESSI flare list\footnote{http://hesperia.nasa.gov/hessidata/dbase/hessi\_flare\_list.txt}, between the start of the RHESSI mission on 2002 February 5 and 2010 January 1.
Accompanying the flare list are ``Quicklook Images'' in different energy bands (3--6, 6--12, 12--25, 25--50, 50--100, and 100--300 keV), typically accumulated over two minutes.
Quicklook Images in a certain energy band are made only if they were deemed to ``reliably image'' the X-ray source, that is,
that an X-ray source was observed to be roughly at the same location in a majority of RHESSI sub-collimators.
We examined all events for which such images above 25 keV existed. There were 838 of them.
\subsection{Derivation of altitude-energy relationship}
We assume that non-thermal electrons precipitate along magnetic field lines that radially extend from the photosphere to the corona.
They propagate from an acceleration region somewhere in the corona, and lose energy and emit bremsstrahlung HXR as per the thick-target model \citep[e.g.][]{Brown2002}.
Under these assumptions, the difference in source altitude, ${\Delta}h_i$, between energy $\varepsilon_i$ and a reference energy $\varepsilon_{ref}$ can be used to approximate the average density between these two points \citep[see e.g.][and Appendix \ref{appendix:directderivation}]{Brown2002}:
The idea is that for an accelerated electron distribution with a negative power-law index propagating towards region of higher densities,
non-thermal emission at energy $\varepsilon$ spatially peaks where electrons have crossed a column density $N \approx \frac{\varepsilon^2}{2K}$, with $K$ a constant.
Hence, knowing the distance $s$ between peaks at emission $\varepsilon_1$ and $\varepsilon_2$, one can get an average density between the two peaks of emission: $n=\frac{N_2-N_1}{s}=\frac{\varepsilon_2^2-\varepsilon_1^2}{2Ks}$.
In actuality, $N = \frac{\varepsilon^2}{2K}$ is only a handy approximation: as discussed in \citet{Brown2002} \citep[see also][]{Xu2008},
there is a dependence on the spectral distribution of injected electrons, and also on the density profile of the medium through which the electron beam is propagating.
While both dependences can easily be taken into account when using forward-fitting techniques on singular events, both vary from flare to flare,
and we have thus decided to use the $N = \frac{\varepsilon^2}{2K}$ relationship throughout our statistical study.
We have computed the RHESSI visibilities \citep{Hurford2002}, accumulated over three minutes around peak HXR flux, in the following energy bands: 6--10, 10--15, 15--20, 20--25, 25--30, 30--35, 35--40, 40--50, 50--60, 60--70, 70--80, 80--90, 90--100.
We have used the hsi\_vis\_fwdfit.pro routine to find positions of centroid at different energies, using subcollimators (SCs) 3--9.
This method basically yields the position of the centroid of the flux (i.e. its center of gravity).
The software also yields error bars on centroid positions.
Typically, higher energies, having less count statistics, yield larger error bars.
We omitted cases where the software did not converge to a solution.
Once the centroid positions are determined at all energy bands and for all flares, we statistically determine the altitude difference, $\Delta{h}_i$,
between emission at all energies $\varepsilon_i$ and a reference energy $\varepsilon_{ref}$.
The $\Delta{h}_i$ are derived using the same method as in \citet{Sato2006} (and described in more mathematical details in Appendix~\ref{appendix:dh}).
Figure~\ref{fig:rvsdr} shows plots of $R$ vs. $\Delta{R}$, with $R$ the projected (on the usual plane of the Sun, perpendicular to the observer's line of sight) distance between HXR emission and Sun center,
and $\Delta{R}$ the projected distance between the centroid of emissions at different energies.
The slope $\frac{\Delta{R}}{R}$ is obtained by linear fitting {\it with a zero intercept coordinate} (i.e. going through the origin).
Because high energies have lesser statistics, leading to poorer positional accuracy of the centroid,
and because there is no potential contamination in the determination of non-thermal centroid position by thermal emission below 35 keV,
(although other effects exist and are discussed in Section~\ref{sect:contamination}), the reference energy was chosen to be 35 keV throughout.
Note that there are many outliers, but that their error bars are generally larger, thus have a lesser contribution to the determination of $\frac{\Delta R}{R}$.
Finally $\Delta{h}_i$=$r_s \frac{\Delta{R}}{R}$, with $r_s$=solar radius (see Appendix~\ref{appendix:dh}).
\begin{figure*}[ht!]
\centering
\includegraphics[width=12cm]{r_vs_dr.eps}
\caption{
{\it Abscissa:} $R$ the projected distance between Sun center and flare emission,
{\it ordinates:} ${\Delta}R$, the projected distance between emission centroid at 35--40 keV and at the energy band indicated on top of each plot, with error bars.
}
\label{fig:rvsdr}
\end{figure*}
From the $\Delta{h}_i$ values obtained in the preceding paragraph, the plot in Figure~\ref{fig:main} (top) is constructed.
\begin{figure}[ht!]
\centering
\includegraphics[width=7.5cm]{mainfigure.ps}
\caption{
{\bf Top:} Plot of $\Delta{h}$=$h-h_{ref}$ as a function of energy $\varepsilon=h\nu$, with $h_{ref}$ being the unknown altitude of peak reference energy $\varepsilon_{ref}$ emission.
$\varepsilon_{ref}$ was chosen to be 35 keV (see text for more details).
The error bars were propagated from the $\frac{\Delta R}{R}$ fittings.
{\it Black:} using all our events, {\it blue:} using all our events with maximum pileup error below 20\%, {\it magenta:} using all our events with maximum pileup error above 20\%.
{\bf Bottom:} Density structure (assuming fully-ionized plasma), as a function of $\Delta{h}$=$h-h_{ref}$ derived from the top plot (see text for details).
Some data points at high densities have no error bars: this means the error bar was actually larger than the nominal value.
We kept them on the plot because of their consistency.
The squares represent the densities derived around 35 keV emission.
}
\label{fig:main}
\end{figure}
It shows the altitude differences between emission at reference energy $\varepsilon_{ref}$=35 keV and emission at other energies (we have also looked at other choices of $\varepsilon_{ref}$, and found similar results, albeit with larger error bars when $\varepsilon_{ref}$ was higher than 35 keV).
Also shown are curves derived from subsets of our data: {\it blue:} using only events with maximum pileup error $<$20\%, and {\it magenta:} using only events with maximum pileup error $>$20\%.
Pileup issues will be discussed further on.
The results presented in Figure~\ref{fig:main} (top) show a systematic decrease in altitude as a function of energy, despite somewhat largish error bars at high energies.
Notice at low energies the departure from a strictly monotonic relationship.
We attribute this to contamination of our sample by thermal emission at low energies.
We have checked the reliability of results in Figure~\ref{fig:main} (top) by varying certain selection criteria and redoing the procedure.
We have tried to select events with quicklook images above 50 keV (as opposed to 25 keV),
tried other subcollimator combinations (e.g. 1--9), tried varying accumulation time intervals, different energy bands, and different $\varepsilon_{ref}$:
the curve essentially remains the same, though generally with larger error bars.
We have also verified that a histogram of the normalized residuals from the $\frac{\Delta{R}}{R}$ linear fitting are normally distributed, with mean 0 and standard deviation 1.
The data presented in Figure~\ref{fig:main} (top) will be used in Section~\ref{sect:Derivation} to determine an average solar chromopheric/coronal density structure during flaring times.
It can already be clearly seen that above $\sim$50 keV, flare footpoint centroids have a vertical extension of less than $\sim$0.5 Mm,
i.e. electrons above $\sim$50 keV reach their stopping heights within that 0.5 Mm region.
We finally discuss the potentially important sources of contamination to our dataset in the next subsection.
\subsection{Possible sources of contamination}\label{sect:contamination}
\begin{itemize}
\item {\it Presence of a thermal component:} At low energies (e.g. usually $\lesssim$15 keV for M-flares, and $\lesssim$25 keV for X-flares),
the non-thermal thick-target beam model can no longer be applied, due to the presence of high-temperature, high-altitude X-ray emitting loops.
\item {\it Presence of a non-thermal coronal source:} \citep{Krucker2008b} These sources are often difficult to observe, due to instrumental dynamic range, but are probably present most of the time \citep{Krucker2008}.
It is likely that such high-altitude source exist, slightly moving the center of gravity of our footpoint sources to higher projected altitudes.
Assuming a coronal source altitude of 10 Mm above the footpoint(s), with 10\% of the footpoint non-thermal flux, the upward shift in altitude of the emission centroid could be $\sim$1 Mm.
As coronal source spectra are much steeper than footpoint spectra \citep[e.g.][]{Marina2006}, this effect is less pronounced at higher energies.
\item {\it Pulse pileup:}
At high photon fluxes, detector pulse-pileup occurs \citep{Smith2002}.
This has the effect of combining two low-energy photon into a single higher-energy photon.
Hence, an intense thermal loop at $\sim$18 keV will have a trace in $\sim$36 keV images (in the case of attenuator state 3).
A pile-up error of 10\% (as given e.g. by the routine hsi\_pileup\_check.pro) under attenuator state 3, should roughly shift the center of gravity of 36 keV emission by $\sim$1 Mm to higher altitudes (assuming the thermal loop is about 10 Mm above the footpoints)
This effect will typically occur only at high countrates, and is very energy-dependent, i.e. present mostly at twice the energy where most of the counts are, itself dependent on the spacecraft attenuator state:
$\sim$12 keV for attenutator state 0, $\sim$24 keV for attenutator state 1, $\sim$36 keV for attenutator state 3.
\item {\it Albedo:}
Albedo effects \citep{Bai1978,Kontar2010} can also influence the results: up to 20--40\% of the 30--40 keV flux can be due to albedo (assuming isotropic beaming).
Assuming most of the 30--40 keV emission comes from $\sim$2 Mm above the photosphere, a $\sim$30\% component reflected from the photosphere would shift the centroid of 30--40 keV emission by about $\sim$0.5 Mm downwards.
This effect is present for all flares, is expected to be greatest around 30--40 keV, and shifts the centroid to lower altitudes, contrary to the pulse-pileup and coronal source effects.
Furthermore, the amplitude of the effect of albedo is maximum near disc center, and minimum near the limb, and hence has a tendency to offset (in part) the effects of pulse pile-up and coronal sources.
\end{itemize}
To conclude, it can be said that observed heights for energies $\gtrsim$50 keV are trustworthy (within their statistical limitations),
but that non-thermal emission $\lesssim$40 keV may be slightly offset (typically by up to 1--2 Mm) to higher altitudes than in reality,
and with the magnitude of this offset varying somewhat with energy
With these caveats in mind, we have attempted in the next section to derive an average density-height profile from our dataset.
\section{Derivation of densities}\label{sect:Derivation}
\subsection{Direct derivation} \label{sect:direct}
Figure~\ref{fig:main} (bottom) displays the density structure obtained using all the flares in our list, and using two subsets of it, with error bars.
They were obtained using the method described in \citet{Brown2002}, and in Appendix~\ref{appendix:directderivation}.
A fully-ionized corona was assumed, but high energy electrons (producing most of the high-energy radiation) probably reach regions of low ionization in the lower chromosphere.
This means that densities obtained at low altitudes (derived from the higher energies) should actually be multiplied by $\sim$2.8 \citep[electron beams in a neutral plasma emits less HXR bremsstrahlung][]{Brown1973,Kontar2002}.
Figure~\ref{fig:main} suggests that
1) emission at lowest energies are clearly contaminated by thermal loops, and results derived from those assuming non-thermal beam models are not to be trusted;
2) emission above 50 keV is emitted from regions with densities $>$10$^{13}$ cm$^{-3}$, assuming fully-ionized plasma.
As we are probably in the chromospheric neutral region, the densities are likely $\sim$3 times as much.
3) emission at intermediate energies ($\sim$20--40 keV) appear to come from regions of different densities, depending on the amount of pileup, itself dependent on the intensity of the non-thermal flux.
From 2), one can hypothesize that $>$50 keV emission always comes from the same altitudes, whether we are dealing with small or large flares.
This in turn tells us that emission at $\sim$35 keV for high pile-up events is almost 1 Mm above that of low pile-up events (top plot of Figure~\ref{fig:main}),
which is in accordance with the discussion in Section~\ref{sect:contamination}, but does not preclude the following scenario:
larger (and longer) flares, which typically have high pileup, are expected to have had greater chromospheric ablation by the time we reach the peak in non-thermal flux (when our observation are made),
making the loop denser, and the emission at intermediate energies ($\sim$35 keV) higher. What we observe is an average of this effect over all our selected flares.
Figure~\ref{fig:main} (bottom) suggests an atmosphere with density profile with at least two exponential components.
In the next section, we have formulated simple density models and attempted to fit it to the data in Figure~\ref{fig:main}(top).
\subsection{Density model fitting} \label{sect:fitting}
We have attempted to fit our data with several density models, but will only discuss a two-exponential with unit step ionization (at altitude $h_{step}$: fully-ionized for altitudes $h>h_{step}$ and fully neutral for $h<h_{step}$) model, and only briefly mention some of the others.
As in \citet{Kontar2008a}, we have added a data point for our fittings, which actually helps us in determining $h_{ref}$, our reference height for emission at $\varepsilon_{ref}$=35 keV:
it is the well-established hydrogen density at the photosphere $n(h=0)=n_0$=1.16$\times$10$^{17}$ cm$^{-3}$.
We have used the routine mpfit.pro (found e.g. in the IDL Astronomy Library) to make our fittings (and tried also IDL's amoeba.pro routine, with no discernible differences).
We have used a Monte Carlo approach to determine error bars for our fitting parameters: for each fitting ``run'' (a hundred such runs were executed), a random amount was added to each data point.
This amount is normally distributed, with mean 0, and standard deviation equal to the nominal error of the data value.
We show only the two-barometric component with a unit step ionization change model (Figure~\ref{fig:fit2exp_step}) for the low pileup case,
because we want to minimize the influence of pulse-pileup.
Data with higher pileup and/or other more complicated models produced large $\chi^2$ results.
We only fitted above $\sim$20 keV. Events with low pileup errors have typically little to no thermal emission above this threshold.
\begin{figure}[ht!]
\centering
\includegraphics[width=8cm]{fit_2exp_step2.ps}
\caption{
Two-barometric component atmosphere fitting on low-pileup data ($<$15\%), with unit step ionization variation.
{\bf Top:} Actual data points and fit (the horizontal ``error bars'' are actually binwidths).
{\bf Bottom:} Fitted density structure {\it (solid black)},
VAL-C atmospheric model \citep[][ {\it solid red}]{VAL1981},
FAL-P atmospheric model \citep[][ {\it solid purple}]{Fontenla1993},
CICM \citep[Caltech Irreference Chromosheric Model][ {\it solid green}]{Ewell1993},
and \citet{Aschwanden2002} {\it (olive green)} results (which assumed a power-law density distribution and a fully-ionized atmosphere).
{\it Solid orange:} FAL-P ionization level, {\it dashed orange:} VAL-C ionization level.
}
\label{fig:fit2exp_step}
\end{figure}
\begin{table*}[ht!]
\caption{Fitting parameters and other parameters derived from them. Double exponential structure, with unit step ionization, using flares with low pileup.}
\centering
\begin{tabular}{ccccc}
\tableline\tableline
Fitting parameters & Description & Units & Value \\
\tableline
$H_0$ & Low-altitude scale height & km & 131$\pm$16 \\
$h_{01}$ & Altitude of transition between exponential components & Mm & 1.69$\pm$0.21 \\
$H_1$ & High-altitude scale height & Mm & 5.4$\pm$0.6 \\
$h_{step}$ & Altitude of unit step ionization change& Mm & 1.3$\pm$0.2 \\
$h_{acc}$ & Altitude of acceleration region & Mm & 19$\pm$32 \\
$\chi^2$ & & - & 0.79 \\
\tableline
\tableline
Derived & & & \\
\tableline
$n_{acc}$ & Density at acceleration altitude& cm$^{-3}$& $(8\pm71)\times10^8$ \\
$\varepsilon_{step}$ & Initial electron energy required to reach neutral layer & keV & 36$\pm$6 \\
\tableline
\tableline
\end{tabular}
\label{tab:2exp_step}
\end{table*}
Table~\ref{tab:2exp_step} summarizes the result of the fittings.
The $h_{step}$ parameter is the altitude where the ionization abruptly changes from 100\% (corona) to 0\% (chromosphere), and $\varepsilon_{step}$ is the minimal initial energy required for electrons to reach the neutral layer.
$h_{01}$ is the altitude of the change from the low-altitude exponential component to high-altitude exponential component.
The scale heights $H_0$ and $H_1$ are trustworthy, thanks to the additional $n_0$ datapoint and to the tight error bars at lower energies.
The acceleration altitude $h_{acc}$ and density $n_{acc}$ are unsurprisingly poorly constrained.
The altitude of ionization change, $h_{step}$, is found to be $\approx$1.3$\pm$0.2 Mm above the photosphere, close to the 5\%--10\% ionization level in the VAL-C or FAL-P atmospheres (Figure~\ref{fig:fit2exp_step}).
The low-altitude scale height was determined to be $\sim$131$\pm$16 km, which corresponds to a (neutral gas) temperature of 5600$\pm$700 K, consistent with chromopheric temperatures below the transition region.
The high-altitude scale height of the flaring atmosphere is determined to be $\sim$5.4 Mm, which, assuming a fully-ionized isothermal plasma, corresponds to a temperature of $\sim$115 kK.
As this region of the lower corona is very dynamic (i.e. non isothermal) during flares, and this value is an average over more than 800 flares, we suspect this usually mid-transition region temperature does not have any intrinsic value.
%
We would like to point out that events with higher maximum pulse pileup error required another exponential component at intermediate altitudes in order to produce a ``reasonable'' fit (albeit with reduced-$\chi^2 >$4).
This led to a fit very close to the CICM and \citet{Aschwanden2002} results at intermediate altitudes (1.5--3 Mm), and, at low and high altitudes, similar to the shown low-pileup case.
While an appealing result, we do not think it is trustworthy, given the afore-mentioned issues.
\section{Conclusion and future work} \label{sect:discussion}
Using most of RHESSI's observations of flares with strong non-thermal HXR emission, we have derived an average emission energy vs. relative height profile.
From this curve, it has been possible to derive densities, assuming a beam-like thick-target model.
Furthermore, adding the well-known density at the surface of the photosphere as an additional data point,
we could derive an average {\it absolute} height vs. density profile, and successfully fitted a two-exponential with step ionization model atmosphere to it.
Although imaging show larger vertical spatial extent of a few Mm \citep{Kontar2010a}, we have found that flare footpoint {\it centroids} above $\sim$50 keV extent vertically, as expected from the thick-target beam model, and are within $\sim$0.5 Mm of each other, in regions with neutral densities well above 3$\times$10$^{13}$ cm$^{-3}$.
We have found density scale heights at low altitudes/high energies of 131$\pm$16 km, matching well the VAL-C or FAL-P models \citep[as well as the 155$\pm$30 km value found by ][]{Kontar2010a}, and of 5.4$\pm$0.6 Mm at higher altitudes, for a flaring atmosphere.
A host of other fitting parameters at intermediate heights or energies, and other derived quantities, could be obtained, but various intrumental, physical and/or statistical effects prevented us from obtaining reliable results for these.
We have listed the different major sources of contamination and gave example of their possible contributions: thermal loops, non-thermal coronal source, pulse pileup, and albedo.
The fitting results we presented minimized the impact of pulse-pileup, and were not affected by thermal contamination.
While properly accounting for all these contaminating effects is difficult in a statistical study (particularly the contribution of any coronal source), we believe they are quantifiable and can be compensated for in a few well-chosen events.
But in keeping with the statistical approach used so far, we plan on investigating the use of backprojection-based imaging with a very limited set of {\it fine} subcollimators:
it is our hope that the lower sensitivity stemming from the use of a smaller number of collimators will be offset by the fact that taking the brightest pixel (as opposed to a centroid) in {\it spatially-resolved} sources should be less prone to the contaminants we have discussed so far.
Our goal is to better resolve the intricate interplay between changes in density and changes in ionization level at intermediate altitudes ($\approx$1--3 Mm above photosphere, corresponding to intermediate energies of $\approx$ 25--50 keV).
|
\section{Introduction}
The star formation activity is fundamental to studies on the formation and evolution of galaxies. Numerous efforts have been made to find reliable and convenient SFR indicators \citep[e.g.][]{kenn98,hira03,bell03,hopk03,calze07}. Among most frequently used indicators, the ultraviolet (UV) and the optical recombination lines (e.g. ${\rm H}\alpha$, ${\rm Pa}\alpha$) give direct measures of light from young stars. However, UV and ${\rm H}\alpha$ emissions are strongly affected by dust extinction \citep{kenn98}, and ${\rm Pa}\alpha$ as well as other optical recombination lines were also shown to underestimate SFRs for galaxies with high luminosity \citep{rieke09}. On the other hand, since dust absorbs UV/optical light and re-emits the bulk of the energy into far-infrared (FIR) band ($25\ \mu{\rm m}\ \sim\ 350\mu{\rm m}$), FIR emission could efficiently trace SFRs for dusty galaxies. However, FIR emission is unable to trace the dust-unobscured radiation and includes part of radiation from old stellar populations. Therefore the energy balance method combining the FIR and UV derived SFRs were used to complement the emission from young stars not traced by FIR \citep[e.g.][]{bx96, meur99, gord00, buat05}. Although this method could trace SFRs with considerable accuracy, it is difficult to obtain the total dust emission especially for high redshift objects.
The MIR monochromatic fluxes were also investigated as SFR indicators. The MIR emission is contributed by several components, including the polycyclic aromatic hydrocarbon (PAH) features (prominent at $6.2$, $7.7$, $8.6$, $11.3$, $12.7$ and $17\ {\rm \mu m}$), the continuum by the stochastic heating of very small grains, silicate absorption at $9.7$ and $18\ {\rm \mu m}$, molecular hydrogen lines and fine-structure lines \citep{lp84, lege89, dese90, dl07, smith07, trey10}. The MIR-SFR relation was intensively studied using Spitzer IRAC $8\ {\rm \mu m}$ and MIPS $24\ {\rm \mu m}$ photometry data and spectral data \citep[e.g.][]{for04, calze05, calze07, kenn09, rieke09, trey10}. Nevertheless, there is still a debate about the reliability of MIR indicators because of the complicated features it contains. Attempts to combine the optical and IR indicators show that this combination could trace the SFR effectively, and notably reduce the scatter \citep[e.g.][]{calze07, zhu08, kenn09}. However, factors of the combination are different in each work.
The recently released AKARI/IRC Point Source Catalogue Version $\beta$-1 (hereafter IRCPSC) provides positions and fluxes of all-sky survey at {\it S9W} ($9\ {\rm \mu m}$) and {\it L18W} ($18\ {\rm \mu m}$) bands \citep{ishi10}. For numerou MIR data, it would be useful if there is a benchmark of SFR measurement. Comparing with Spitzer $8$ and $24\ {\rm \mu m}$ bands, the AKARI IRC {\it S9W} and {\it L18W} bands cover wider wavelength ranges, including silicate absorption features at both bands and emission contributed by large PAH molecules at the {\it L18W} band. This work is dedicated to investigate whether and to what degree AKARI broadband MIR data could trace SFRs, and how much the inclusion of silicate absorption and longer wavelength PAH features would be.
Start from a sample with multi-wavelength observation flux, we derive the SFR for each galaxy by the spectral energy distribution (SED) fitting. MIR data were correlated with the SFRs to built the SFR calibrations. Then the results were compared with those from the Spitzer observation. The present paper is organized as follows: Section 2 introduces the GALEX-SDSS-2MASS-AKARI sample. Section 3 gives a short introduction to methods for calculating SFRs. The MIR-SFR relations and some comparisons between AKARI and Spitzer are reported and discussed in Section 4. Conclusions are given in Sections 5.
\section{Data}
\subsection{Construction of the multi-wavelength sample}
We cross-identified\footnote{The searching radius is set at $36''$ considering the original sample's position accuracy.} IRCPSC with multi-wavelength data constructed by \citet{take10} to obtain a sample including MIR photometric measurements. The original sample was based on a selection by IRAS-PSC${\it z}$ \citep{saun00} and AKARI/FIS Bright Source Catalog (hereafter FISBSC) data \citep{yama08, yama09}, which means all these galaxies have considerable fluxes at FIR. Then these galaxies were cross-matched with observations by GALEX, 2MASS and SDSS. Fluxes at the UV band were obtained by performing specific aperture photometry so as not to shred GALEX images. A detailed description of the original sample can be found in \citet{take10}.
IRCPSC presents flux data at $9\ \mu{\rm m}$ and $18\ \mu{\rm m}$ that have an effective bandwidth of $4.10\ \mu{\rm m}$ and a detection limit of $50\ {\rm mJy}$, and $9.97\ \mu{\rm m}$ and $90\ {\rm mJy}$, respectively \citep{ishi10}. After cross-identification, there were 162 galaxies with a flux of either $9\ \mu{\rm m}$ or $18\ \mu{\rm m}$. However, nine galaxies were found to have inconsistent fluxes at GALEX, SDSS or 2MASS band according to their SEDs, possibly due to the measurement errors or the misidentification of objects caused by the inhomogeneous resolution of each observation; they were omitted from our sample. We also discarded one galaxy with a too-small redshift ($z\ \sim\ 0.0008$). The summary of the sample is listed in Table \ref{tab:data_sum}. Searching in the SIMBAD database, most of galaxies in our sample are normal star forming galaxies: Only 15 galaxies are classified as AGNs (including Seyfert 1 and Seyfert 2 galaxies). All the galaxies in our sample are nearby ones. The distribution of the redshift is shown in Figure \ref{fig:z_distri}.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_z_distri.ps}
\end{center}
\caption{The redshift distribution of our sample. The filled area denotes AGNs.}\label{fig:z_distri}
\end{figure}
\subsection{The Re-estimation of AKARI FIR Data}
The FISBSC flux density for extended sources would be no longer accurate because of the point source extraction procedure. In our sample, a considerable fraction of galaxies is the extended source. Therefore, it is questionable whether the catalog data of these sources are reliable. This is confirmed by making a comparison between FISBSC fluxes and IRAS co-added fluxes (Figure \ref{fig:iras_cat}), which were specially calculated for extended sources \citep{saun00}. In order to investigate the dependence of FISBSC flux on the extension of the galaxy, the sample was divided into three sub-samples according to the size of the galaxy (Table \ref{tab:div}). The lines in Figure \ref{fig:iras_cat} give predictions of the difference between two different bands by \citet{dh02} (hereafter DH) one-parameter constrained SED templates. The parameter $\alpha$ is related to IRAS flux ratio $f_{60}/f_{100}$. The $\log(f_{60}/f_{100})$ in our sample ranges from $-0.60$ to $0.25$, and most of the galaxies have $\log(f_{60}/f_{100})$ between $-0.5$ and $0.0$, corresponding to an $\alpha$ value between $2.625$ and $1.375$. Accordingly, the model prediction with $\alpha\ =\ 1.375$, $2.625$ and the median value $2.0$ are illustrated in Figure \ref{fig:iras_cat}. Table \ref{tab:data_comp} presents the median values of $\log(f_{\rm AKARI}/f_{\rm IRAS})$ for each sub-sample. Compared with the model predicted values, there are clear discrepancies between FISBSC and IRAS data, especially for $60-65\ \mu{\rm m}$ and $90-100\ \mu{\rm m}$. Also note that the large galaxies show greater discrepancies than small ones, indicating the PSF photometry is less reliable for larger galaxies, because a considerable part of their flux is left out by the relatively small beam size.
\begin{figure*}
\begin{center}
\FigureFile(150mm,150mm){bw_4bands_iras_dh_cat.ps}
\end{center}
\caption{Comparison between FIS catalog flux and the IRAS flux. Galaxies in different sub-sample are shown in different sizes (large open circles: large; small open circles: medium; dots: small). The DH models with three different $\alpha$ are shown in different lines (dashed:$\alpha=1.375$; solid: $\alpha=2.0$; dash-dotted: $\alpha=2.625$).}\label{fig:iras_cat}
\end{figure*}
Aiming to obtain more reliable flux data, the photometry of the diffuse maps provided by the AKARI group was conducted using SExtractor. The obtained ``AUTO'' fluxes were also compared with IRAS co-added fluxes.
In the $90\ \mu{\rm m}$ band, the result shows a great improvement on the consistency with the IRAS flux according to the DH model (Figure \ref{fig:iras_map} and Table \ref{tab:data_comp}) for all galaxies in our sample. At $65$ and $140\ \mu{\rm m}$ bands, the consistency is improved for those galaxies in `Medium' and `Large' subsamples, whereas the dispersion increases for the galaxies in 'Small' samples. Less improvement is found at $140\ \mu{\rm m}$ band, which can be explained by the relatively larger PSF FWHM at this band \citep[$\sim$ $60''$, as compared with $\sim$ $39''$ for the $65$ and $90\ \mu{\rm m}$ bands,][]{kawa07}. At the $160\ \mu{\rm m}$ band, neither fluxes derived from the diffuse maps nor those from the IRCPSC are satisfying, which is due to the poor quality of this band, thus $160\ \mu{\rm m}$ data were not used in the following work. Therefore, we kept IRCPSC flux values for `Small' sources at the $60$ and $140\ \mu{\rm m}$ bands, and applied fluxes derived from diffuse maps to the other sources. In addition, fluxes derived from AKARI diffuse maps showed notably smaller measurement errors than the IRAS fluxes, greatly improving the data quality. Considering taht the IRAS fluxes provide information similar to that given by the AKARI bands, we omitted IRAS fluxes when fitting the SEDs.
\begin{figure*}
\begin{center}
\FigureFile(150mm,150mm){bw_4bands_iras_dh_map.ps}
\end{center}
\caption{Comparison between FIS diffuse map flux and the IRAS flux. Symbols are of the same meaning as in Figure \ref{fig:iras_cat}}\label{fig:iras_map}
\end{figure*}
Note that the MIR data provided by IRCPSC were little affected by an extension of the source, since it applies ``AUTO'' fluxes by SExtractor, which are suitable for both point sources and extended sources.
\section{SFR Calculation}
The SED fitting program CIGALE \citep{noll09} was used to calculate the SFR for our sample. CIGALE was developed to derive highly reliable galaxy properties by fitting the UV/optical SEDs and the related dust emission at the same time, i.e., the stellar population synthesized models are connected with infrared templates by the balance of the energy of dust emission and absorption. A detailed description of CIGALE can be found in \citet{noll09, buat10, giov10}. Here we give a brief introduction to the main features.
CIGALE allows one to use the stellar SEDs from models given either by \citet{mara05} or by \citet[][PEGASE]{fr97}. The difference between these models is the contribution of the thermally pulsating asymptotic giant branch (TP-AGB) stars. In PEGASE models, the contribution from TP-AGB stars is low \citep{mara06}. \citet{mara05} increased the contribution from TP-AGB stars adopting the 'fuel consumption' approach. \citet{mara06} shows that the insufficient consideration of TP-AGB stars overestimates the stellar mass by 0.2 dex and worsens the consistency with IR observation data at a redshift of $z\ \sim\ 2$. At lower redshift, it is found that using different models little affects the results \citep{rutt06, emin08}. Therefore, the models from \citet{mara05} were preferred in this work. The Kroupa IMF \citep{krou01} was used to calculate the complex stellar populations (CSPs).
This code provides two scenarios of star formation: One is ``box models'' with a constant SFR, the other is ``$\tau$ models'', for which SFR decreases exponentially with a typical decay time, $\tau$. SFR is calculated as ${\rm SFR}_{\rm box}\ =\ M_{\rm gal}/t$ for ``box models'' and ${\rm SFR}_{\tau}\ =\ M_{\rm gal}/[\tau(e^{t/\tau}\ -\ 1)]$ for ``$\tau$ models'', where $M_{\rm gal}$ is the galaxy mass \citep{noll09}. CIGALE also allows one to apply different scenarios for young and old populations. The input SFH here is a constant burst SFH for young stellar populations, and an exponentially decreasing one for old stellar populations. Thus SFRs were calculated using the formula $f_{\rm ySP}\ \cdot\ {\rm SFR}_{box}\ +\ f_{\rm oSP}\ \cdot\ {\rm SFR}_{\tau}$, where $f_{\rm ySP}$ and $f_{\rm oSP}$ are fractions of young stellar populations and old stellar populations, respectively.
The attenuation curve adopted by CIGALE is based on a law given by \citet{calze00}, with modification of the slope and/or adding a UV bump. The modification of the slope is controlled by the factor $(\lambda/\lambda_{V})^{\delta}$, i.e., by changing $\delta$, the slope of the attenuation curve can be modified. We only considered a modification of the slope of the attenuation law here, and no bump was introduced. CIGALE allows one to consider different effects of attenuation for old and young stellar populations by adding the reduction factor $f_{att}$ of the dust attenuation for the old stellar populations as an input parameter. For the IR part, CIGALE uses DH models, which is described in Section 2. Then the dust emission was calculated, and by balancing the energy emitted and absorbed, the short and long wavelength parts of the model were connected.
To compute the output parameters the code provides two methods: ``sum'' and ``max''. The former calculates the probability distribution functions (PDFs) by taking sums of the probability of the models in given bins of parameter space, which might cause an unintentional bias when the input parameter values are badly chosen. The latter introduces a fixed number of equally sized bins for each parameter and searches the maximum probability of models in each bin. Then these maximum probabilities are taken as weights for individual bins to calculate the expectation value of the parameter. The advantage of this method is that it alleviates the dependence on the choice of parameters \citep[see][for detail]{noll09}. Therefore, we applied the ``max'' method in this work.
The input parameters applied in this study were adopted from \citet{buat10} in order to obtain a stable and reliable output. Since CIGALE is unable to trace the unobscured emission of an AGN, for Seyfert 1 galaxies, the output decreases the reliability \citep{buat10}. For dust-obscured AGNs, CIGALE provides models to fit the SED, and can avoid introducing any severe bias. Therefore, five Seyfert 1 galaxies were rejected in our sample, and other AGNs are marked during analysis. In order to keep the accuracy of derived SFRs, galaxies with relatively large discrepancies between observation and output spectra (reduced $\chi^2\ >\ 10$) were discarded as being unreliable. At first running, there were 25 such galaxies. By updating the SDSS data using Navigator of SDSS DR7/8 instead of the original pipeline data from SDSS DR7, the reduced $\chi^2$ values of 17 galaxies decreased to less than 10. \footnote{There is no modification for other sources which are fitted well.} Four of the eight remaining galaxies had very poor quality of AKARI diffuse maps. The other four had large extension and brightness, which could cause incomplete flux derivation or saturation in the optical plates. Considering that the number was small ($\sim$ 6\%), these eight galaxies and five out of the 17 galaxies were discarded. Therefore, reliable SFRs were derived for 140 galaxies, out of which there are 112 with available $9\ \rm{\mu m}$ fluxes and 97 with $18\ \rm{\mu m}$ fluxes.
\subsection{The reliability of the results}
For a SED fitting, the accuracy of the output depends on the input parameters. To give proper estimates of the SFRs, the robustness of the results must be tested. A straightforward way to check the reliability of the output of CIGALE is to use a sample of mock galaxies with comprehensively known physical parameters. Mock galaxies can be generated following a recipe in \citet{giov10}: 1, Run CIGALE on the data of real galaxies. For each galaxies, a best model is produced by $\chi^2$ minimization. Then, from these models, the fluxes at each bands can be estimated. 2, Add to each flux a random relative error, which is normally distributed with $\sigma\ =\ 0.1$. Thus, we obtain a mock catalog with flux information at every photometric band used in this study. The last step is to run the code on the mock catalog and then to compare the output parameters with the exact values provided by the best models. The result of the comparison is shown in Figure \ref{fig:mock}. We only present the result concerning SFRs because in this work the SFR is the only parameter that needs to be of concern. Figure \ref{fig:mock} shows the two quantities are well related, indicating SFRs derived here are reliable.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){mock_check_irc.ps}
\end{center}
\caption{The comparison between SFRs derived from CIGALE code and from the mock galaxies.}\label{fig:mock}
\end{figure}
As discussed in \citet{noll09}, CIGALE could provide stable results of SFRs as long as one constraint beyond PAH band is given. We ran CIGALE with and without MIR data to examine the influence on SFR by MIR photometric data. The result shows adding the MIR data or not adding has almost no influence on the resulting value of SFR (Figure \ref{fig:cmp_cig}), while not using FISBSC data brings large uncertainties to the output, consistent with the conclusion of \citet{noll09} using SINGS sample.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_cmp_cig.ps}
\end{center}
\caption{The comparison between SFR derived with and without MIR data. AGNs are shown as triangles.}\label{fig:cmp_cig}
\end{figure}
\section{Result and Discussion}
Apparently, the luminosity at $9\ \rm{\mu m}$ and $18\ \rm{\mu m}$ $L_{9}$ and $L_{18}$ \footnote{In this paper, $L_{\lambda}$ refers to $\nu L_{\nu}$ at wavelength $\lambda$.} both correlate with SFRs (Figure \ref{fig:res09} and Figure \ref{fig:res18}); the Spearman's correlation coefficients are 0.943 for the $L_{9}$-SFR relation and 0.956 for the $L_{18}$-SFR relation\footnote{Pearson's correlation coefficient is 0.944 for $L_{9}$-SFR relation and 0.951 for $L_{18}$-SFR relation}. Linear regressions following the method provided by \citet{kell07} give:
\begin{equation}
\log\ \frac{{\rm SFR}}{M_{\odot}/{\rm yr}}\ =\ (0.99\pm 0.03)\log\ \frac{L_{9}}{L_{\odot}} - (9.02 \pm 0.32)
\label{equ:res09}
\end{equation}
and
\begin{equation}
\log\ \frac{{\rm SFR}}{M_{\odot}/{\rm yr}}\ =\ (0.90\pm 0.03)\log\ \frac{L_{18}}{L_{\odot}} - (8.03 \pm 0.30).
\label{equ:res18}
\end{equation}
The scatters of the data points about the regression lines of Equations \ref{equ:res09} and \ref{equ:res18} are approximately the same, with $\sigma\ =\ 0.18$ dex for $9\ \mu{\rm m}$ and $0.20$ dex for $18\ \mu{\rm m}$. These tight correlations also hold for the surface densities of luminosities and SFRs (Figures \ref{fig:res09_sf} and \ref{fig:res18_sf}). The correlation coefficients are 0.961 for $9\ \mu{\rm m}$ and 0.945 for $18\ \mu{\rm m}$. The areas of galaxies were calculated from {\it g}-band images of SDSS. The regression gives:
\begin{eqnarray}
\nonumber \log\ \frac{\Sigma_{\rm SFR}}{M_{\odot}{\rm yr}^{-1}{\rm kpc}^{-2}}\ &=&\ (1.02\pm 0.03)\log\ \frac{\Sigma_{9}}{L_{\odot}{\rm kpc}^{-2}} \\
&-&\ (9.30 \pm 0.19)
\label{equ:res09_sf}
\end{eqnarray}
with $\sigma\ =\ 0.18$, and
\begin{eqnarray}
\log\ \frac{\Sigma_{\rm SFR}}{M_{\odot}{\rm yr}^{-1}{\rm kpc}^{-2}}\ &=&\ (0.98\pm 0.04)\log\ \frac{\Sigma_{18}}{L_{\odot}{\rm kpc}^{-2}} \\
&-&\ (8.89 \pm 0.25).
\label{equ:res18_sf}
\end{eqnarray}
with $0.22$.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_res09_re.ps}
\end{center}
\caption{The $9\ {\rm \mu m}$ luminosity-SFR relation. The dashed line shows the fitting result. The triangles are AGNs.}\label{fig:res09}
\end{figure}
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_res18_re.ps}
\end{center}
\caption{The $18\ {\rm \mu m}$ luminosity-SFR relation. Lines and symbols share the same meaning with Figure \ref{fig:res09}.}\label{fig:res18}
\end{figure}
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_res09_sf.ps}
\end{center}
\caption{The surface densities of $9\ {\rm \mu m}$ luminosity-SFR relation. The dashed line shows the fitting result. The triangles are AGNs.}\label{fig:res09_sf}
\end{figure}
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_res18_sf.ps}
\end{center}
\caption{The surface densities of $18\ {\rm \mu m}$ luminosity-SFR relation. Lines and symbols share the same meaning with Figure \ref{fig:res09}.}\label{fig:res18_sf}
\end{figure}
Figures \ref{fig:res09} and \ref{fig:res18} show that AGNs share a MIR-SFR relation similar to normal star-forming galaxies, although the radiation mechanism of AGNs is different from normal galaxies. The slopes of the $\log L_{9}$-$\log$SFR and $\log L_{18}$-$\log$SFR relations derived here are almost equal to one, indicating the MIR-SFR relations are close to linear.
\subsection{Comparison with SFR calibrations from Spitzer data}
Spitzer $8\ \rm{\mu m}$ and $24\ \rm{\mu m}$ data were investigated as SFR tracers by several authors \citep[e.g][]{wu05, Perez06, calze07, relano07, zhu08, rieke09}. The $24\ \rm{\mu m}$ fluxes were found to be tightly related to the emission of the warm dust, and thus more intensively investigated, whereas the $8\ \rm{\mu m}$-SFR relation is more complicated, strongly depending on such as the metallicity, size and star-formation history \citep{calze07}, therefore, there are fewer calibrations.
The main difference between the AKARI and Spitzer filters is their bandwidths. Due to the wider band, AKARI $9\ \rm{\mu m}$ and $18\ \rm{\mu m}$ fluxes are more affected by silicate absorption, PAH and molecular hydrogen line emissions (Figure \ref{fig:filter_res}). In order to check the reliability of our calibrations, the SFRs derived from Equations \ref{equ:res09} and \ref{equ:res18} were compared with those given by Spitzer calibrations. The work to compare was chosen to keep the luminosity range close to the present sample (Table \ref{tab:sfr_cal}). The calibrations given by \citet{wu05} and \citet{zhu08} are based on an equation given by \citet{kenn98} in which Salpeter IMF was used. The use of Salpeter IMF will cause smaller SFRs by $\sim\ 0.18$ dex than some other IMF with a more shallow slope at low masses \citep{rieke09}. Therefore this effect was corrected for the results of \citet{wu05} and \citet{zhu08}.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){filter_res.ps}
\end{center}
\caption{The filter response curves of AKARI $9\ \rm{\mu m}$ and $18\ \rm{\mu m}$ bands (dashed line) and Spitzer $8\ \rm{\mu m}$ and $24\ \rm{\mu m}$ bands (dotted line). The solid line is the luminosity weighted average spectrum of star forming galaxies from \citet{smith07}.}\label{fig:filter_res}
\end{figure}
\subsubsection{Comparison between SFRs derived from $9\ \rm{\mu m}$ and from $8\ \rm{\mu m}$}
Spitzer $8\ \rm{\mu m}$ fluxes were computed from the output spectra of the CIGALE and filter response curves of Spitzer IRAC and MIPS. Since the $8\ \rm{\mu m}$ fluxes used in the calibration given in Table \ref{tab:sfr_cal} were the dust emission with the stellar contribution subtracted following the recipe of \citet{helo04}, the $3.6\ \mu{\rm m}$ flux was also calculated to compute the stellar composition contained in $8\ \rm{\mu m}$, and thus $8\ \rm{\mu m}$ dust emission could be obtained (hereafter, we refer to $8\ \rm{\mu m}$ dust emission as $8\ \rm{\mu m}$ emission for conciseness). Note that here the stellar contribution is very small, which can only affect the result by $\sim\ 0.02$ dex. The obtained flux is then converted to SFR by formula given by previous work (Table \ref{tab:sfr_cal}). The results are plotted in Figure \ref{fig:comp8_9} (The typical $1\sigma$ uncertainty for the galaxy of median luminosity is $\sim$ 0.5 dex). The statistical information of the comparison is given in Table 5.
The discrepancy between our results and \citet{wu05} may due to several reasons. \citet{wu05} listed factors such as the accuracy of fiber aperture corrections, the validity of the estimation of the obscuration in galaxies by using Balmer decrement, the possible contamination to radio and MIR emission from obscured weak AGNs. The larger capacity of our sample (79 for $8\ {\rm \mu m}$ in \citet{wu05} compared with 112 in our sample) and the wider coverage of $9\ {\rm \mu m}$ band may also cause such difference. Another possible reason is that the oversimplification of PAH emission in DH models underestimates the $8\ {\rm \mu m}$ flux and therefore gives smaller SFRs. However, this level of discrepancy is well within the scatters in Equations \ref{equ:res09}.
The discrepancy between \citet{wu05} and \citet{zhu08} is because \citet{zhu08} included $8\ {\rm \mu m}$-weak H${\rm II}$ galaxies with lower MIR luminosity \citep{zhu08}. Since no such galaxies were included in present sample, it is reasonable that present result agrees with \citet{wu05} better.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){compare8_9_re.ps}
\end{center}
\caption{Comparison between the SFRs derived from $9\ {\rm \mu m}$ emission (Equation \ref{equ:res09}) and from $8\ {\rm \mu m}$ emission by \citet{wu05} (crosses) and \citet{zhu08} (circles).}\label{fig:comp8_9}
\end{figure}
\subsubsection{Comparison between SFRs derived from $18\ \rm{\mu m}$ and $24\ \rm{\mu m}$}
For a $18\ {\rm \mu m}${--}$24\ {\rm \mu m}$ SFRs comparison, more reference calibrations are available (Table \ref{tab:sfr_cal}). The $24\ {\rm \mu m}$ fluxes are also derived from the output spectra of the SED fitting by CIGALE. The statistical information of the comparison is given in Table \ref{tab:18_24}. The converted SFRs are plotted in Figure \ref{fig:comp18_24}. Since the SFRs derived from \citet{wu05} are quite similar to those from \citet{zhu08} and the luminosity range in \citet{wu05} is closer to this work, the results from \citet{zhu08} are omitted for the sake of brevity.
The results of \citet{wu05} and \citet{zhu08} agree well with our result after the correction of IMF. \citet{rieke09} assembled SED templates for local luminous and ultraluminous infrared galaxies and combined the result of \citet{dale07} and \citet{smith07} to produce templates at lower luminosities. Their result is applicable to galaxies with $24\ {\rm \mu m}$ luminosity greater than $6\ \times\ 10^8\ L_{\odot}$, corresponding to $\log\ {\rm SFR}\ =\ -0.33$, which is shown by the dotted line in Figure \ref{fig:comp18_24}. The present result agrees very well with \citet{rieke09} above the limit.
Our result is a little higher than the one given by \citet{calze07}. A possible reason is that the result of \citet{calze07} was derived for H${\rm II}$ clouds by Pa$\alpha$ emission, which might be poorly applied to galaxy-wide calculations, because the diffuse MIR or Pa$\alpha$ emissions in the whole galaxy are not included \citep{alon06, calze07, kenn07, rieke09}.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){compare18_24_re.ps}
\end{center}
\caption{Comparison between the SFRs derived from $18\ {\rm \mu m}$ (Equation \ref{equ:res18}) and from $24\ {\rm \mu m}$ emission by \citet{wu05} (crosses), \citet{calze07} (triangles) and \citet{rieke09} (dots). The dotted line gives the lower limit where the calibration of \citet{rieke09} applies.}\label{fig:comp18_24}
\end{figure}
\subsection{Combination of FUV and MIR indicators}
At lower IR luminosity, the IR indicator may fail to trace part of the UV photons from young stars due to the increased transparency of the ISM. A combination of unobscured FUV and MIR luminosities, (${\rm FUV}\ +\ \alpha{\rm MIR}$), may efficiently compensate for the lost energy, and could trace the SFR linearly \citep{zhu08}. However, upon converting SFRs to dust obscuration corrected FUV fluxes by eq.[1] from \citet{kenn98}, we find that $\alpha\ =\ 2.53$ and $3.33$ with a scatter of $0.17$ dex for $9\ \rm{\mu m}$ and $0.20$ dex for $18\ \rm{\mu m}$, respectively. The scatter is not reduced significantly. This fact indicates that the origin of the scatter in MIR-SFR diagram is complicated: not only the untraced UV photons, but also other unknown factors, such as the variation of the physical conditions within each galaxy, the distribution of dust and photo dissociation regions (PDRs), etc.
\subsection{Metallicity}
We attempt to investigate the gas-phase metallicity range of our sample by searching in the metallicity database measured by \citet{trem04} for SDSS galaxies. Unfortunately, the $12\ +\ \log\rm{(O/H)}$ values are given for only 33 galaxies (all higher than 8.75). Therefore, we applied a compromised method: to investigate the stellar mass $M_{*}$ of our sample. Measured by CIGALE, the stellar mass $M_{*}$ of all the galaxy in the sample is larger than $10^{8.5}$ $M_{\odot}$. Thus, from mass-metallicity relation given by \citet{trem04}:
\begin{eqnarray
\nonumber 12\ +\ \log({\rm O/H})\ =\ &-&\ 1.492\ +\ 1.847(\log M_{*}) \\
&-&\ 0.08026(\log M_{*})^2,
\end{eqnarray}
the metallicity $12\ +\ \log\rm{(O/H)}$ for our sample is higher than $8.4$. This is not surprising because the initial sample from \citet{take10} is IR selected, which means a considerably high luminosity in IR and thus sufficient dust content and relatively high metallicity range. Thus, the MIR-SFR relations derived here could only be extrapolated to other high-metallicity galaxies. The situation for low metallicity galaxies is rather complicated; since the opacity of the galaxy decreases, MIR would be unable to trace most of the UV/optical photons and thus lose the ability as an SFR indicator \citep{calze07}.
\subsection{AGNs}
Although AGNs have distinct features from normal galaxies in various physical properties, they share the same trend in the MIR-SFR figures. A possible reason is that the contribution from AGN component is minor (from SED fitting, less than 15\%), therefore the host galaxy component dominates the spectrum. To investigate the effect of AGNs on MIR emission, we plot average SEDs for AGNs and normal galaxies in Figure \ref{fig:ave_sed}. On average, AGNs are brighter than normal galaxies at all bands, while with a lower MIR/$90\ {\rm \mu m}$ ratio (Table \ref{tab:mir_90}). There are two possible reasons: 1, silicate absorption occurs more strongly in AGNs. 2, PAH molecules are destroyed by the harsh radiation field in AGNs. The latter is more convincing. Studies show that small PAH molecules contributed to shorter wavelength MIR emission are destroyed more easily than large ones contributed to longer wavelength MIR emission \citep[][and reference therein]{smith07, trey10}, which is consistent with the lower $9\ {\rm \mu m}$ flux value than $18\ {\rm \mu m}$ in Figure \ref{fig:ave_sed}. Nevertheless, these differences between AGNs and normal galaxies are only on their average level, which could not be used to distinguish AGNs and normal galaxies in individual cases.
\begin{figure}
\begin{center}
\FigureFile(80mm,80mm){bw_ave_sed.ps}
\end{center}
\caption{The average SED (dashed lines) and the SED normalized at $90\ {\rm \mu m}$ (solid lines) of AGNs (solid symbols) and normal galaxies (open symbols).}\label{fig:ave_sed}
\end{figure}
\section{Conclusion}
We combined AKARI/IRC $9\ {\rm \mu m}$ and $18\ {\rm \mu m}$ data with a previous sample to construct a FIR selected multi-wavelength sample with MIR photometric measurements. The FIR data of AKARI/FIS in the original sample were re-estimated by photometry of AKARI diffuse maps to correct the bias of PSF photometry for extended sources. Then, the SEDs of the sample were fitted by CIGALE, and the SFRs were obtained. Regression analysis was conducted to investigate MIR-SFR relations. SFRs converted from AKARI MIR fluxes were compared with those from the Spitzer MIR fluxes to test the reliability of AKARI MIR-SFR calibrations. From the result, we draw the following conclusions:
\begin{enumerate}
\item Both $9\ {\rm \mu m}$ and $18\ {\rm \mu m}$ luminosities correlate with SFRs, and thus could be converted SFRs.
\item A combination of FUV and MIR luminosity barely reduces the scatters, indicating that the unobscured UV photons are not the only reason of the variation of MIR-SFR relation.
\item A comparison of the SFRs derived from Equations \ref{equ:res09} and \ref{equ:res18} with the ones derived from Spitzer MIR-SFR relations shows that the silicate absorption included in {\it S9W} ($9\ {\rm \mu m}$) and {\it L18W} ($18\ {\rm \mu m}$) bands little affects the results. The discrepancies, if any, are well within the uncertainties.
\item AGNs in the sample show no discrepancy with normal galaxies in the MIR-SFR diagrams. The smaller average MIR fluxes for AGNs than normal galaxies might indicate the small PAH molecules are destructed by harsh radiation from AGNs.
\end{enumerate}
In summary, for IR selected galaxies the rest frame $9\ {\rm \mu m}$ and $18\ {\rm \mu m}$ emissions are efficient tracers of SFRs, and the equations derived here should be applicable to other dust rich galaxies.
\bigskip
This work is based on observations with AKARI, a JAXA project with the participation of ESA. FTY, TTT, and KLM are partially supported from the Grand-in-Aid for the Global COE Program ``Quest for Fundamental Principles in the Universe: from Particles to the Solar System and the Cosmos'' from the Ministry of Education, Culture, Sports, Science and Technology (MEXT). TTT has been supported by Program for Improvement of Research Environment for Young Researchers from Special Coordination Funds for Promoting Science and Technology, and the Grant-in-Aid for the Scientific Research Fund (20740105) commissioned by the MEXT of Japan. VB and DB have been supported by the Centre National des Etudes Spatiales (CNES) and the Programme National Galaxies (PNG). JIP is supported by the grant AYA2007-67965-C03-02, from the Spanish MICINN. This work makes use of the SIMBAD database, operated at the Centre de Donées astronomiques de Strasbourg (CDS), Strasbourg, France. We acknowledge Y. Doi for preparing diffuse maps for our sample and Y. Matsuoka for helpful advice and comments. We also thank the anonymous referee for constructive suggestions to improve this paper.
|
\section{Introduction}
The harvesting of sunlight by organisms such as plants and bacteria plays a crucial role in life on earth. Nature has optimized the photosynthesis process to a great extent, leading to a high efficiency in both the capturing of light and the subsequent energy transport within the organism's light harvesting complexes.~\cite{Amerongen00, Sundstrom99, Sumi99, Mukai99, McDermott95} Pigment molecules in these light harvesting structures absorb the light, after which the excitation energy is eventually funneled towards the reaction center, where charge separation takes place and the energy can be utilized for biochemical purposes. A closer understanding of the functioning of these biological systems and what features are crucial to their high efficiency may provide further insight in how to optimize synthetic light harvesting structures.
Recent experimental studies have shown that quantum coherence in light harvesting complexes persists on surprisingly long timescales.~\cite{Brixner05,Engel07, Lee07, Panit10} While excitation energy transport (EET) in such molecular systems has been a subject of study for decades,~\cite{Forster48, Grover70, Kenkre74} the aforementioned experimental observation of long-lived coherence despite its occurrence in noisy environments at room temperature has spurred further theoretical research. Naively, one might think that coupling the electronic excitations to a dissipative environment will lead to rapid decoherence and a loss of excitation energy. However, it has been shown that environment-induced dephasing and noise may actually enhance the transport efficiency of excitations in such structures.~\cite{Gaab04, Mohseni08, Plenio08, Ishizaki09, Rebentrost09, Wu10} The rationale behind this lies in the fact that interaction with degrees of freedom in the bath can induce transitions of the excitation that are directed towards the reaction center. Recent studies have shed further light on the mechanisms that allow for optimal efficiency of the excitation, suggesting an intricate functioning where aspects such as the non-Markovian nature of the bath interactions~\cite{Rebentrost09-2, Roden09, Chen11} and correlations in the bath~\cite{Adolphs06, Caruso09, Wu10, Strumpfer11} may be exploited to achieve further optimization of the light harvesting complex.
The interaction of the chromophores with their local environments leads to fluctuations of the transition energies and interactions. More specifically, changes in the environment will induce fluctuating local electrical fields. The resultant Stark shifts in the transition energies and transition dipoles of the chromophores will thus also fluctuate; in addition, environment-induced changes in relative position or orientation of the chromophores will also cause corresponding fluctuations in the interactions.~\cite{Adolphs06} A number of studies have been performed where only the effect of transition energy fluctuations has been included, as well as possible spatial and temporal correlations between these transition energy fluctuations.~\cite{Adolphs06, Caruso09, Wu10, Strumpfer11} This is commonly done in the framework of the extended Haken-Strobl-Reineker model,~\cite{Haken72, Haken73} which provides an attractive approach due to its simplicity and tractability. The original Haken-Strobl-Reineker model allows for a treatment of both transition energy and interaction fluctuations, and while it does not include correlations and inherently assumes a high temperature, the formalism can straightforwardly be generalized to remove the former drawback. Fluctuations in the interactions, in contrast to transition energy fluctuations, have typically not been accounted for in studies of photosynthetic complexes, despite the fact that these will generally occur as well.
Fluctuations in transition energies and interactions are induced by variations in the local environments of the chromophores, and environmental variations around one chromophore will generate both fluctuations in its transition energy and in its interaction with the other chromophores. In addition, different chromophores may share part of the environment; both these arguments imply that the generated fluctuations will in general be correlated. While correlations between transition energy fluctuations have been studied,~\cite{Adolphs06, Caruso09, Wu10, Strumpfer11} there should generally also be correlations of transition energy fluctuations and interaction fluctuations, and between different interaction fluctuations. In previous studies, such correlations have commonly been discarded, even though there is no a priori reason to do so. A number of recent studies have also suggested the occurrence of such fluctuations in photosynthetic complexes.~\cite{Ulrich11, Shim11, Coker11} Moreover, fluctuations of the various intermolecular interactions are usually not accounted for, although also these are typically induced by changes in the local environment. We present a study of the effects of fluctuations of the interactions and their possible correlations with the various transition energy and other interaction fluctuations, and provide a numerical analysis of their measurable effects. When applied to natural photosynthetic complexes, such as the Fenna-Matthews-Olson (FMO) complex,~\cite{Fenna75, Olson80, Blankenship88, Li97, Vulto98} it is shown that interaction fluctuations can have a considerable impact on the excitation dynamics and the efficiency, and these effects should generally not be neglected. Furthermore, we show that additional correlations can lead to enhanced oscillations of the exciton populations and modifications of the transfer and dephasing rates, and we quantify the dependence of the trapping efficiency on the correlations and initial conditions.
The paper is structured as follows. In Sec. \ref{Sec:Theory}, we provide the theoretical background of our study: in Sec. \ref{Sec:HSR}, we introduce the Haken-Strobl-Reineker formalism that is used to describe the effect of environmentally induced fluctuations and their possible correlations on the excitation dynamics, Sec. \ref{Sec:corr} discusses the possible correlations, and in Sec. \ref{Sec:FMO} we introduce the Fenna-Matthews-Olson photosynthetic complex that we will apply our theory to. In Sec. \ref{Sec: numerical}, we show the results of application of our theory to FMO: Sec. \ref{Sec:uncorrelated fluctuations} elucidates the role of uncorrelated interaction fluctuations, Sec. \ref{Sec:EJfluc} focuses on the effect of correlations between fluctuations in transition energies and interactions, while subsequently in Sec. \ref{Sec:JJfluc} we investigate the effect of correlated interaction fluctuations. Our conclusions are presented in Sec. \ref{Sec:conclusions}.
\section{Theory}
\label{Sec:Theory}
\subsection{The Haken-Strobl-Reineker model}\label{Sec:HSR}
Upon absorption of the solar light and energy transfer into the FMO complex, an electronic excited state will be created in the bacteriochlorophylls (BChl's). Due to the strong interchromophore interactions, this excited state (exciton) will be delocalized over a number of BChl's. To describe the excitation and its dynamics, we employ the Frenkel exciton Hamiltonian,~\cite{Davydov71, Agranovich82}
\begin{equation}H=H_S+H_R+H_{S-R}=\sum_nE_n \left|n\right>\left<n\right|+\sum_{n,m\neq n} J_{nm}\left|n\right>\left<m\right|+\sum_{q}\omega_qb_q^{\dagger}b_q+H_{S-R},\end{equation}
where $\left|n\right>$ describes a state where chromophore $n$ is in its excited state while all others are in their ground states, $J_{nm}$ is the interaction between chromophores $n$ and $m$, and we have a bath of harmonic modes labeled by $q$. The first two terms are the system Hamiltonian $H_S$, the third term corresponds to the bath Hamiltonian $H_R$, and the final term describes the coupling between the system and the bath $H_{S-R}$, which we will keep unspecified for now. In our case, the system consists of the BChl's within the photosynthetic complex, while the environment includes any other degrees of freedom the excitations can couple to, such as the protein surroundings. To adequately describe the environmental effects, it is necessary to work in the density matrix formalism.~\cite{Neumann27, Blum81, May00} The density matrix corresponding to the wave function $\left|\psi(t)\right>$ is given by $\tilde{\rho}(t)=\left|\psi(t)\right>\left<\psi(t)\right|$, and evolves according to the Liouville-von Neumann equation ($\hbar$ has been set to unity), \begin{equation}\label{Liouville} i \frac{\partial \tilde{\rho}(t)}{\partial t}=\left[H(t),\tilde{\rho}(t)\right]\equiv \mathcal{L}\tilde{\rho}(t).\end{equation}
Here, we have defined the Liouville superoperator as $\mathcal{L}=\left[H,...\right]$. The above time evolution equation of the density matrix is equivalent to the time evolution of the wave function $\left|\psi(t)\right>$ as given by the Schr\"odinger equation. At this point, we switch to the interaction picture, where the time evolution induced by the Hamiltonian terms $H_S+H_R$, which are large compared to the system-bath interaction $H_{S-R}$, is explicitly removed from the time evolution of all operators $A(t)$, \begin{equation}A_I(t)=e^{i\left(H_S+H_R\right)t}A(t)e^{-i\left(H_S+H_R\right)t}.\end{equation}
We are primarily interested in the evolution of the system degrees of freedom, which we obtain by taking the trace over the bath degrees of freedom. This leaves us with the reduced density matrix $\rho(t)$ in the interaction picture, which is the quantity we will consider from this point on when referring to the density matrix.
The approach we take in treating interactions with the environment is based on a method first introduced by Haken, Strobl and Reineker.\cite{Haken72, Haken73} In this approach, one models the bath-induced fluctuations as classical Gaussian Markov processes. While the original Haken-Strobl-Reineker (HSR) model assumed uncorrelated site energy fluctuations, one can extend the methodology to allow for correlations between the various fluctuations; we refer to Ref. \onlinecite{Chen10} for the details. The Hamiltonian can then be written as \begin{equation}\label{HamHSR}H=\sum_nE_n \left|n\right>\left<n\right|+\sum_{n,m\neq n} J_{nm}\left|n\right>\left<m\right|+\sum_{n,m}V_{nm}(t)\left|n\right>\left<m\right|.\end{equation} The terms $V_{nm}(t)$ are the environmentally induced fluctuations of the various system Hamiltonian matrix elements. Note that the Hermiticity of the Hamiltonian implies $V_{nm}(t)=V_{mn}^{*}(t)$. The averages $\left<V_{nm}(t)\right>$ can be set to zero without loss of generality, and since the fluctuations are assumed to be Gaussian Markov processes, the problem is fully defined when the correlation functions $\left<V_{nm}(t)V_{n'm'}(t')\right>$ are known. The effect of the environment is now fully encoded in the bath correlation functions $C_{nmn'm'}(\tau)=\left<V_{nm}(\tau)V_{n'm'}(0)\right>$.
\subsection{Correlation functions}\label{Sec:corr}
To proceed, we make the white noise assumption, where it is assumed that the bath relaxes on a time scale that is short compared to the exciton dynamics. The time dependence of the bath correlation functions is taken as a $\delta$-function,
\begin{equation}C_{nmn'm'}(\tau)\equiv
\left<V_{nm}(\tau)V_{n'm'}(0)\right>=\gamma_{nmn'm'}\delta(\tau).\end{equation}
Substitution into the time evolution equation of the density matrix, Eq. \ref{Liouville}, yields
\begin{equation}\label{timeevolution}\frac{d\rho_{nm}(t)}{dt}=-i\mathcal{L}_{sys}\rho_{nm}(t)+\sum_{n'm'}\left[\gamma_{nn'm'm}\rho_{n'm'}(t)+\gamma_{m'mnn'}\rho_{n'm'}(t)-\gamma_{nn'n'm'}\rho_{m'm}(t)-\gamma_{n'mm'n'}\rho_{nm'}(t)\right].\end{equation}
At this stage, it is worthwhile to consider the correlation matrix $\hat{\gamma}$ in more detail. First of all, the elements $\gamma_{nnnn}$ and $\gamma_{nmnm}$($n\neq m$) are simply the variances of respectively the transition energy fluctuations and the interaction fluctuations. All other elements of $\hat{\gamma}$ correspond to correlations between different fluctuations. Secondly, the correlation matrix fulfills a number of symmetries, due to the requirement of a Hermitian Hamiltonian and a set of trivial index permutation invariances,
\begin{equation}\gamma_{abcd}=\gamma_{cdab}=\gamma_{bacd}=\gamma_{abdc}.\end{equation}
We observe that there are generally three types of correlations, namely between fluctuations in \begin{itemize} \item excitation energies ($a=b,c=d$), \item excitation energies and interactions ($a=b, c\neq d$ or vice versa), \item interactions ($a\neq b,c\neq d$).\end{itemize}
A number of studies have included the first type of correlation, i.e. spatially correlated transition energy fluctuations, which have been shown to lead to appreciable changes in the transfer rates and efficiency of light-harvesting systems.~\cite{Adolphs06, Caruso09, Wu10, Strumpfer11} However, the second type of correlation has only been studied in the dimer,~\cite{Chen10} while the third type of correlation has to the best of our knowledge not been included at all in studies of EET in photosynthetic complexes. The central concern of this study is to provide an understanding of the relevance of such correlations to the time evolution of the various density matrix elements and the overall quantum efficiency.
Besides the above symmetry considerations, additional restrictions apply to the various correlations in order to describe a physical system.~\cite{Chen10} This becomes clear when one considers the expectation value of the difference between two fluctuation elements squared, which should obviously give a positive result, \begin{equation}\left<\left(V_{nm}-V_{n'm'}\right)^2\right>\propto \gamma_{nmnm}+\gamma_{n'm'n'm'}-2\gamma_{nmn'm'}\geq0.\end{equation} By considering such inequalities for various values of the indices, one finds a number of consistency conditions that need to be fulfilled,
\begin{align}\label{consistency}
\gamma_{nmnm}\geq 0\\
2\left|\gamma_{nmnm'}\right|\leq\gamma_{nmnm}+\gamma_{nm'nm'},\end{align}
where the indices $n$,$m$ and $m'$ are allowed to be equal. In addition, the magnitude of the cross-correlation between two fluctuating matrix elements is limited by the magnitude of the original fluctuations. It is straightforward to show that \begin{equation}\gamma_{nmn'm'}\leq \sqrt{\gamma_{nmnm}\gamma_{n'm'n'm'}},\end{equation} where the equality holds when the fluctuations of the Hamiltonian matrix elements $V_{nm}$ and $V_{n'm'}$ are fully (anti-)correlated. These limitations on the allowed fluctuation correlations should be kept in mind when considering the effect of including the various additional correlations.
\subsection{The Fenna-Matthews-Olson complex and light harvesting efficiency}\label{Sec:FMO}
The model light harvesting system for which we want to quantitatively probe the effects of correlations in the environmentally induced fluctuations is the Fenna-Matthews-Olson (FMO) complex. FMO is the photosynthetic complex of the green sulfur bacterium Chlorobium Tepidum, consisting of three weakly coupled, identical subunits of seven bacteriochlorophyll-a (BChl) molecules each.\cite{Ermler94, Li97, Vulto98, Cho05, Brixner05, Adolphs06, Engel07, Lee07, Panit10} Within such a subunit, after absorption of the incoming light, the excitation energy is transferred to the reaction center where charge separation can take place. Due to their proximity to the primary light harvesting antennae, energy typically flows into the system at BChl's 1 and 6.~\cite{Li97}
We model a subunit of the FMO complex by the following experimental Hamiltonian in the site basis (in units of cm$^{-1}$),\cite{Vulto98, Cho05}
\begin{equation}\label{hamFMO} H= \left(\begin{array}{ccccccc}
280 & -106 & 8 & -5 & 6 & -8 & -4 \\
-106 & 420 & 28 & 6 & 2 & 13 & 1 \\
8 & 28 & 0 & -62 & -1 & -9 & 17 \\
-5 & 6 & -62 & 175 & -70 & -19 & -57 \\
6 & 2 & -1 & -70 & 320 & 40 & -2 \\
-8 & 13 & -9 & -19 & 40 & 360 & 32 \\
-4 & 1 & 17 & -57 & -2 & 32 & 260
\end{array} \right). \end{equation}
The corresponding exciton states, which we label in order of increasing energy, are given in Appendix \ref{appendixA}. Of particular relevance for our numerical results are the exciton states $s=3$ and $s=7$ with energies of respectively $E_3=224$ cm$^{-1}$ and $E_7=480$ cm$^{-1}$, which are the exciton states mostly associated with BChl's 1 and 2, and the lowest energy exciton state $s=1$ with energy $E_1=-24$ cm$^{-1}$ which is localized mostly on BChl 3, from where energy will be trapped to the reaction center.
In order to describe the influence of the various correlations on the light harvesting efficiency, we introduce two competing decay channels.~\cite{Ritz01, Cao08, Wu10} Exciton decay, leading to an irreversible loss of the absorbed energy, is described by adding a decay term \begin{equation}\mathcal{L}^{(dec)}_{nm}=\left(k^{(d)}_n+k^{(d)}_m\right)/2,\end{equation} where $k^{(d)}_n$ is the exciton decay rate at chromophore $n$. In the FMO complex, energy is transferred to the reaction center, and we model capture of the excitation by the trap through
a trapping term \begin{equation}\mathcal{L}^{(trap)}_{nm}=\left(k^{(t)}_n+k^{(t)}_m\right)/2,\end{equation} where $k^{(t)}_n$ is the exciton trapping rate at chromophore $n$. In the FMO complex, the reaction center is located close to BChl 3, and we will only include trapping from that particular chromophore, $k^{(t)}_n=k^{(t)}\delta_{n,3}$.
The efficiency of the process can be defined as the branching ratio between energy trapped at the reaction center and energy lost through exciton decay,~\cite{Ritz01, Cao08, Wu10}
\begin{equation} q= \frac{\sum_n k^{(t)}_n\tau_n}{\sum_n k^{(t)}_n\tau_n+\sum_n k^{(d)}_n\tau_n},
\end{equation}
where $\tau_n=\int_0^{\infty}dt \rho_{nn}(t)$ is the mean residence time at site $n$.
\section{Numerical results for the FMO complex}\label{Sec: numerical}
As stated before, the FMO complex consists of seven coupled BChl's. In the following sections, we will first discuss the effects of uncorrelated interaction fluctuations, and subsequently proceed with a study of the effect of correlations between the various possible fluctuations. Since many of the interactions are already small to begin with, we only consider fluctuations in the strongest interactions. Likewise, we only consider correlations involving the strongest interactions, i.e. interactions between BChl's of a magnitude exceeding $25$ cm$^{-1}$, which are shown in Fig. \ref{fig:fmoschema}. In addition, we only include correlations between site energy fluctuations of site $n$ and interactions involving that same site, and similarly, correlations between interaction fluctuations that have one site in common. This choice is motivated by the fact that correlations between fluctuations are predominantly caused by shared local environments.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.5\columnwidth]{fmo3.eps}
\end{center}
\caption{A schematic view of the FMO complex, with the strongest inter-BChl interactions denoted by arrows. The excitation will typically enter the complex at either BChl 1 or BChl 6, and will flow towards the reaction center which is most closely associated with BChl 3.
}
\label{fig:fmoschema}
\end{figure}
The numerical calculations reported in the upcoming sections concern energy flowing into the FMO complex at BChl 1. As we label the exciton states $s$ in order of increasing energy, this corresponds to initially exciting exciton states $s=3$ and $s=7$. One of our goals is to investigate the effect of including energy-interaction fluctuation correlations on the experimentally observed oscillations of the exciton populations.\cite{Engel07} Therefore, we choose an initial excitation where exciton coherence is already initially present, $\left|\psi(0)\right>=\left(\left|s=3\right>+e^{i\theta}\left|s=7\right> \right)/\sqrt{2}$. Obviously, different values of the mixing angle $\theta$ correspond to different initial populations $P_n$ of BChl's 1 and 2. In particular, $\theta=0$ implies $P_1=0.08$ and $P_2=0.89$, $\theta=\pi/2$ yields $P_1\approx P_2=0.49$, and for $\theta=\pi$ we have $P_1=0.90$ and $P_2=0.08$.
Additional calculations have been performed for energy flowing in from the other side, that is, from BChl 6. This corresponds (mostly) to initially exciting exciton states $s=5$ and $s=6$. The results do not change qualitatively, except for a considerable decrease in population oscillation frequency due to the smaller energy difference between exciton states $s=5$ and $s=6$ as compared to exciton states $s=3$ and $s=7$. For future reference, the exciton that most strongly overlaps with the trap state BChl 3 is exciton state $s=1$.
The parameters considered in our simulations have previously been used in Ref. \onlinecite{Wu10} to model the FMO system. We take all exciton decay rates equal at a value $k^{(d)}_n=k^{(d)}=1$ ns$^{-1}$, while only trapping from BChl 3 occurs, with a trapping rate $k^{(t)}=1$ ps$^{-1}$. Furthermore, we take $\gamma_0=\gamma_{nnnn}=94$ cm$^{-1}$ for the transition energy fluctuations; note that there is a factor of 2 difference in the definition of $\gamma_{nnnn}$ as compared to the quantity $\Gamma$ in Ref. \onlinecite{Wu10}.
\subsection{Uncorrelated interaction fluctuations} \label{Sec:uncorrelated fluctuations}
The effect of interaction fluctuations can be investigated straightforwardly with the current formalism, by introducing nonzero values for the fluctuations of the off-diagonal Hamiltonian matrix elements, $\gamma_{nmnm}\equiv \gamma_1$. As stated in the previous section, only fluctuations of the strongest interactions, shown in Fig. \ref{fig:fmoschema}, are included. Note that, since $\gamma_{nmnm}$ is the expectation value of a squared quantity, it is necessarily positive. In the original works by Haken, Strobl and Reineker,\cite{Haken72, Haken73} interaction fluctuations were already considered, and also a recent paper by Chen and Silbey~\cite{Chen10} showed that interaction fluctuations can have a pronounced effect on exciton dynamics in the dimer. This makes it all the more surprising that this effect is commonly neglected in recent studies concerning the FMO complex. The time evolution of the population of exciton state $s=3$ and the (real part of the) coherence between exciton states $s=3$ and $s=7$ is shown in Fig. \ref{fig:gamma1}, clearly showing that their time evolution strongly depends on the presence of interaction fluctuations.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur2.eps}
\end{center}
\caption{Time evolution of the exciton population $\rho_{33}$ (solid lines) and the real part of the exciton coherence $\rho_{37}$ (dashed lines), with initial phase $\theta=0$, for $\gamma_0=94$ cm$^{-1}$ and $\gamma_1=0,1,2,5,10,20$ cm$^{-1}$. Note the appreciable increase in both transfer rate and decoherence rate with $\gamma_1$.
}
\label{fig:gamma1}
\end{figure}
Analysis of the temporal evolution of the exciton populations and coherences, in particular the ones shown in Fig. \ref{fig:gamma1} and estimating the rates by assuming simple exponential decay and looking at times where the oscillations have died out, reveals that the population transfer rate and decoherence rate both increase approximately linearly with the interaction fluctuations $\gamma_1$. This is to be expected from the general form of the time evolution equations, Eq. \ref{timeevolution}. While these full equations contain a very large amount of terms, all couplings between the density matrix elements are proportional to either $\gamma_0$ or $\gamma_1$. The population transfer rate, in particular, contains only a small contribution from $\gamma_0$ terms for these parameters, and therefore the population transfer is increased dramatically upon introduction of interaction fluctuations. Note that this behavior is consistent with previous results for the dimer,~\cite{Wertheimer82, Chen10} where the exciton population transfer rate for a strongly detuned dimer is to a very large extent dominated by the interaction fluctuations. This is analogous to the situation in the FMO complex, where the differences between the BChl transition energies are also typically large compared to the interactions. More specifically, for a dimer detuned by an energy difference $\Delta E$ and having an interaction strength $J$, the mixing angle $\theta^{\prime}$ defined by $\tan \theta^{\prime}=2J/\Delta E$ is small. It can analytically be shown~\cite{Wertheimer82, Chen10} that the population transfer rate $\Gamma$ between the two exciton states is given by $\Gamma=\gamma_0 \sin^2\theta^{\prime}+2\gamma_1\cos^2\theta^{\prime}$. However, since the first term is typically very small for strongly detuned dimers (i.e., $\Delta E \gg J$), the population transfer rate is fully dominated by the second term as soon as interaction fluctuations are present. An analysis of the population evolution shown in Fig. \ref{fig:gamma1} confirms that the population transfer rates in FMO are indeed also proportional to $\gamma_1$. Thus, interaction fluctuations lead to dramatic increases in population transfer rates, in turn enhancing the efficiency of the FMO complex.
To further illustrate the increase in population transfer rates with increasing interaction fluctuations, Fig. \ref{fig:gamma1-efficiency} shows the harvesting time as a function of the interaction fluctuation amplitude $\gamma_1$. In Fig. \ref{fig:gamma1-efficiency}, we consider the energy fluctuation magnitude $\gamma_0$ that optimizes the harvesting time, $\gamma_0=94$ cm$^{-1}$, and an increase respectively decrease of one order of magnitude. Note that increasing $\gamma_1$ leads to more efficient light harvesting due to a corresponding increase in population transfer within the FMO complex; this holds for different values of the energy fluctuations $\gamma_0$ and for different initial conditions. Additionally, while the optimal dephasing value $\gamma_0$ does not change with $\gamma_1$, the optimum becomes considerably less deep, making the FMO complex more robust to variations in the dephasing rate $\gamma_0$. For very large, in fact unphysically large, interaction fluctuations, the efficiency of the light harvesting complex becomes independent of the energy fluctuation magnitude. This situation corresponds to large population transfer rates, dominated by the interaction fluctuations $\gamma_1$, and subsequent rapid redistribution of population over all BChl's. The harvesting time will in that case converge to $\tau=7$ ps, corresponding to a trapping rate $k_t=1$ ps$^{-1}$ from BChl 3 where at all times 1/7th of the total untrapped population will reside.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur3.eps}
\end{center}
\caption{Dependence of the harvesting time on the interaction fluctuation magnitude $\gamma_1$, for various values of the energy fluctuations $\gamma_0$. Notice the decrease in harvesting time with $\gamma_1$ as a result of increasing population transfer rates.
}
\label{fig:gamma1-efficiency}
\end{figure}
\subsection{Effect of correlated energy-interaction fluctuations} \label{Sec:EJfluc}
\subsubsection{Parameter set 1} \label{subsec: ps1}
Here, we consider $\gamma_0=\gamma_{nnnn}=94$ cm$^{-1}$ which corresponds to parameters close to the optimal energy transport conditions reported in Ref. \onlinecite{Wu10}. The magnitude of the allowed correlations is limited by Eqs. \ref{consistency}; to allow for appreciable amounts of correlation, we take a large interaction fluctuation value $\gamma_1=\gamma_{nmnm}=40$ cm$^{-1}$ ($n \neq m$), with the same provisions as before. In Sec. \ref{subsec: ps2}, we discuss the effect of correlated energy and interaction fluctuations for a smaller, more realistic value of the interaction fluctuation magnitude $\gamma_1$. As stated in Sec. \ref{Sec: numerical}, we only include correlations that involve the strongest interactions (see Fig. \ref{fig:fmoschema}) and we focus on correlations between interaction fluctuations and energy fluctuations of the BChl's involved in that particular interaction; $\gamma_{n'n'nm}\propto\left(\delta_{n'n}+\delta_{n'm}\right)$, $n \neq m$.
First of all, taking all energy-interaction fluctuation correlations equal in sign and magnitude leads to negligible effects. This is consistent with a previous study by Chen and Silbey, where it was shown that in a dimer the relevant quantity is the difference between the two transition energy-interaction fluctuation correlations.\cite{Chen10} While the full time evolution equations for the more complicated FMO system are more involved, a similar result is observed to hold. This is not surprising, as fluctuation correlations of equal sign would imply that a change of interaction would mean that both energies involved change in the same direction, leaving the energy difference unchanged. Therefore, we choose the various energy-interaction correlations of equal magnitude, but with alternating sign; that is, we take $\gamma_{1112}=-\gamma_{2212}=\gamma_{2223}=-\gamma_{3323}\equiv \gamma_2$ and so on. There is an ambiguity near BChl 4, as it strongly interacts with three other BChl's and it is therefore necessary to choose two out of $\gamma_{4434}$, $\gamma_{4454}$ and $\gamma_{4474}$ with equal sign. We take $\gamma_{4434}=\gamma_{4474}=-\gamma_{4454}=-\gamma_2$; other choices result only in inconsequential numerical changes.
A few observations can be made regarding these plots. First of all, energy-interaction fluctuation correlations affect both the energy transfer rates and the coherence decay rates. Depending on the sign of the correlations, both increases and decreases of the various rates can occur. This can be observed in Fig. \ref{fig:33EJ}, where the population of exciton state $s=3$ shows appreciable differences in transfer rate to other exciton states for different correlation magnitudes. While this particular exciton state and initial condition shows an increase in the transfer rate upon increasing correlations, this is not general; a decrease in transfer rate may be obtained for different exciton states and initial conditions. Secondly, it is clear that these parameters lead to a quick redistribution of population over all chromophores, only showing oscillatory features on a short time scale. Therefore, in the next subsection, we consider the same quantities for different parameters, where the effect on population oscillations can be observed more clearly.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur4.eps}
\end{center}
\caption{Time evolution of the exciton population $\rho_{33}$, $\theta=0$, with $\gamma_0=94$ cm$^{-1}$ and $\gamma_1=40$ cm$^{-1}$. The considered correlation values are $\gamma_2=-40,-20,0,20,40$ cm$^{-1}$.
}
\label{fig:33EJ}
\end{figure}
It is also possible to investigate the dependence of the harvesting efficiency, here quantified through the average trapping time $\tau$, on the magnitude of the energy-interaction fluctuation correlation $\gamma_2$. This is shown in Fig. \ref{fig:eff1}; the effects are rather limited, with changes of at most a few percent. Both the overall effect of accounting for these correlations and the dependence on the initial condition, defined by the mixing angle $\theta$, is rather weak. This is related to the fact that we are close to the conditions that constitute optimal harvesting behavior (i.e., $\gamma_0=94$ cm$^{-1}$), which is corroborated by the harvesting times which are only slightly above $\tau=7$ ps. In other words, the bottleneck in the overall harvesting process is not population transfer between the BChl's, but the trapping of excitations from BChl 3 into the reaction center. Therefore, the effect of limited changes in the population transfer rates have relatively little effect on the overall efficiency. Note that negative values of $\gamma_2$ may give a small enhancement of the harvesting time; depending on the initial conditions, a nonzero optimal choice for the correlations may be made such that light harvesting is at its fastest.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur5.eps}
\end{center}
\caption{Dependence of the harvesting time on the magnitude of the energy-interaction fluctuation correlations, for various initial conditions and with $\gamma_0=94$ cm$^{-1}$ and $\gamma_1=40$ cm$^{-1}$. Note that negative values of $\gamma_2$ are required for enhancement of the efficiency, and that the optimal correlation value depends on the initial condition.
}
\label{fig:eff1}
\end{figure}
Again, we note that the fluctuations considered here lead to a rapid decay of the coherences. In the next section, we will consider parameters that lead to slower decoherence and longer lived population oscillations, in order to evaluate the effect of energy-interaction fluctuation correlations on population oscillations.
\subsubsection{Parameter set 2} \label{subsec: ps2}
In order to clearly observe the effect of additional energy-interaction fluctuation correlations on population oscillations, it is necessary to switch to parameters where the decoherence occurs on a slower timescale. Here, we consider $\gamma_0=\gamma_{nnnn}=10$ cm$^{-1}$, and $\gamma_1=\gamma_{nmnm}=2$ cm$^{-1}$ ($n \neq m$) for the interactions shown in Fig. \ref{fig:fmoschema}. The same observations as before hold with regards to the signs of the correlations: alternating signs are required for observable effects and are used throughout this section, while a negative $\gamma_2$ can produce enhancement of the transport efficiency. Again, as an example we consider the time evolution of the population of the exciton states $s=3$ and initial phase $\theta=0$ for various values of $\gamma_2$, shown in Fig. \ref{fig:ps2}.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur6.eps}
\end{center}
\caption{Time evolution of the exciton population $\rho_{33}$, $\theta=0$ and with $\gamma_0=10$ cm$^{-1}$ and $\gamma_1=2$ cm$^{-1}$. We consider the values $\gamma_2=-4,-2,0,2,4$ cm$^{-1}$.
}
\label{fig:ps2}
\end{figure}
It can now be clearly seen that the populations oscillate in time, an effect which is induced by the coherence that is present between the two exciton states. Indeed, the period of the oscillations corresponds to the energy difference between the exciton states, and is identical to the coherence oscillation period, which corresponds to $\tau_{osc}=130$ fs. A clear augmentation of the amplitude of the oscillations can be observed when one increases the magnitude of $\gamma_2$. The introduction of energy-interaction fluctuation correlations can thus enhance the population oscillations, which are driven by the presence of coherence. As before, we also observe a change in population transfer rates, with in particular a strong suppression of population transfer with increasing $\gamma_2$.
The dependence of the harvesting efficiency on the correlation magnitudes can be calculated exactly as before, and is shown in Fig. \ref{fig:eff2}. There is now a stronger dependence on the magnitude of the correlations, which is caused by the fact that this parameter set constitutes suboptimal conditions, so that in this case changes in the population transfer rates have a larger effect on the overall harvesting time. Also, the behavior is almost independent of the initial conditions, even more so than for the parameters in the previous section. Negative values for $\gamma_2$ are required for enhancement, where the transport is now optimized by choosing $\gamma_2$ as negative as possible. Positive values of $\gamma_2$ lead to a strong increase in harvesting time; this corresponds to the strong decrease in the relevant population transfer rates, as observed in for example Fig. \ref{fig:ps2}.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur7.eps}
\end{center}
\caption{Dependence of the harvesting time on the magnitude of the energy-interaction fluctuation correlations, for various initial conditions and with $\gamma_0=10$ cm$^{-1}$ and $\gamma_1=2$ cm$^{-1}$. Note that negative values of $\gamma_2$ are required for enhancement of the efficiency.
}
\label{fig:eff2}
\end{figure}
\subsection{Effect of correlated interaction-interaction fluctuations}\label{Sec:JJfluc}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.8\columnwidth]{Figuur8.eps}
\end{center}
\caption{Time evolution of the exciton population $\rho_{33}$, $\theta=0$, $\gamma_0=94$ cm$^{-1}$, $\gamma_1=40$ cm$^{-1}$, and $\gamma_2=40$ cm$^{-1}$. The considered correlation values are $\gamma_3=-30,-15,0,15,30$ cm$^{-1}$.
}
\label{fig:33JJ}
\end{figure}
It is straightforward to include correlations between the various interaction fluctuations. As in the previous section, we anticipate that such correlations are strongest between interactions that have one molecule in common, and in addition we focus on the strongest interactions. First of all, it is important to note that while such correlations may very well be present, also these are naturally limited in magnitude in order to fulfill the consistency conditions Eqs. \ref{consistency}. In particular, we have
\begin{equation}
\left<\left(V_{ab}-V_{ac}\right)^2\right>\propto \gamma_{abab}+\gamma_{acac}-2\gamma_{abac}>0,
\end{equation}
where $a \neq b, a\neq c$. From this, it follows that
\begin{equation}
\left|\gamma_{abac}\right| \leq \frac{1}{2}\left(\gamma_{abab}+\gamma_{acac}\right).\end{equation}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=0.7\columnwidth]{Figuur9.eps}
\end{center}
\caption{Dependence of the harvesting time on the magnitude $\gamma_3$ of the interaction-interaction fluctuation correlations, for various initial conditions. We have chosen $\gamma_0=94$ cm$^{-1}$, $\gamma_1=40$ cm$^{-1}$ and $\gamma_2$=40 cm$^{-1}$. The different offsets of the curves are consistent with the observed $\theta$-dependence of the harvesting time in Fig. \ref{fig:eff1}.
}
\label{fig:qC}
\end{figure}
Nevertheless, it is instructive to quantify the effects of these additional correlations. The conclusions are typically similar to those in the previous section, although the effects are typically less pronounced. For brevity, we will not show all the same figures as before, but we show a typical result in Fig. \ref{fig:33JJ}. Here, we again consider the parameters previously used in Sec. \ref{subsec: ps1}, with in addition an energy-interaction fluctuation correlation magnitude $\gamma_2=40$ cm$^{-1}$ with the sign conventions defined there. First of all, here too alternating signs are required to amplify the effect of interaction-interaction fluctuation correlations. We take the interaction-interaction fluctuation correlations of the form $\gamma_{1223}=-\gamma_{2334}=\gamma_{3445}=\gamma_3$ et cetera. Again there is some ambiguity in the definition of the signs for the correlations around BChl 4, however, since the details of those choices hardly matter for the initial conditions we have chosen, we will not go in to this in more detail. We show time evolution plots only for $\theta=0$, behavior for other parameters is not significantly different. First of all, it is clear that interaction-interaction fluctuation correlations, too, lead to changes in the population transfer and dephasing rates, as exemplified by the population evolution of exciton state 3 shown in Fig. \ref{fig:33JJ}. In addition, while there is a small change in the size of the populations, the amplitude of the oscillations is influenced by the interaction-interaction fluctuation correlation magnitude to only a small extent. These conclusions are observed to hold for different sets of parameters and initial conditions.
Fig. \ref{fig:qC} shows the dependence of the harvesting time on the magnitude of the interaction-interaction fluctuation correlations. Again, there is an optimum for some nonzero value of the correlation magnitude $\gamma_3$; the exact value of the optimal magnitude for the correlations depends on the initial conditions and on the choice for the energy-interaction correlations. The behavior for other values of the initial phase factor $\theta$ is very similar, implying that the transfer rates and other relevant parameters for trapping will only depend weakly on possible interaction-interaction fluctuation correlations. While Fig. \ref{fig:qC} shows that the effect of interaction-interaction fluctuation correlations is roughly of the same magnitude as the effect of transition energy-interaction fluctuation correlations, this is only the case if the former occurs in conjunction with the latter. The changes induced by interaction-interaction fluctuation correlations for the case of $\gamma_2=0$ cm$^{-1}$ are considerably smaller. Finally, as was previously the case in Sec. \ref{Sec:EJfluc}, a more suboptimal choice of parameters leads to a larger effect of changes in transfer rates, and will thus also amplify the effect of including interaction-interaction fluctuation correlations to an extent.
\section{Conclusions}\label{Sec:conclusions}
Since in principle environmental changes will induce fluctuations of both transition energies and interactions that should be correlated to some extent, we have performed a study of the effect of such fluctuations and their possible correlations on the energy transfer and excitation dynamics in photosynthetic complexes, specifically focusing on the Fenna-Matthews-Olson complex as a model system. The inclusion of interaction fluctuations and correlations between transition energy fluctuations and interaction fluctuations, and between different interaction fluctuations, has been studied by a straightforward generalization of the Haken-Strobl-Reineker model. These additional correlations have been shown to be naturally limited in magnitude, in order to correspond to a physically consistent system.
Interaction fluctuations on the excitation dynamics in photosynthetic complexes can in general not be neglected. Not only will these in general occur, but it has been shown that even small values of interaction fluctuations can already lead to an appreciable increase in transfer rates and a corresponding increase in efficiency of the FMO complex. In particular, for suboptimal values of the energy fluctuations, interaction fluctuations can lead to a considerable optimization of the light harvesting process, even to such an extent that the transfer rates and the eventual efficiency is dominated by interaction fluctuations and not by energy fluctuations. The presence of interaction fluctuations may thus make the efficiency of the FMO complex more robust to changes in the dephasing rate. Generally, for various initial conditions and energy fluctuation magnitudes, an increase in the efficiency of the FMO complex is observed with increasing interaction fluctuation amplitude.
When including correlated transition energy fluctuations and interaction fluctuations, it is first of all possible to observe somewhat enhanced population oscillations, driven by the coherence between the initially excited exciton states. Depending on the signs of the various correlations, a suppression of the oscillations can also occur. In addition, one sees that the transfer rates between exciton states is modified, which in turn leads to accompanying changes in the overall efficiency of the photosynthetic complex. The overall effects on the efficiency are limited in scope when one considers parameter values that approximately correspond to an optimal functioning of the light harvesting complex. In that case, not the excitation transfer but the trapping into the reaction center is the bottleneck, so that changes in transfer rates will only change the harvesting time by up to a few percent. For less optimal conditions, the effects can be considerably larger, as the excitation transfer processes play a larger role in determining the overall harvesting time. Typically, a small increase in the light harvesting efficiency can be obtained by using a nonzero amount of correlation between the aforementioned fluctuations. By the same approach, one can quantify how correlations between the various interaction fluctuations modify previously obtained results. The effect of these additional correlations turns out to be qualitatively very similar to that of transition energy-interaction fluctuation correlations, and are quantitatively of a comparable magnitude. The net effect of such correlations is limited to small changes in the scattering rate and resultant changes in the overall efficiency.
This study thus suggests that one should not neglect interaction fluctuations in models describing the energy transfer in photosynthetic complexes. Correlations may also play a relevant role in the excitation dynamics, and this may (but does not necessarily) translate into appreciable changes in the efficiency.
\section*{Acknowledgments}
The authors thank Dr. Jianlan Wu, Dr. Xin Chen and Dr. Jianshu Cao for useful discussions. This work is supported by ARO under grant W911NF-09-0480, and DARPA grant N66001-10-1-4063.
|
\section{\bf Introduction}
\label{intro}
The average velocity law near the wall is well known from many years (\cite{Karman30}) for what concerns the flow in smooth pipes (\cite{Nikuradse}, \cite{Reichardt}) and in the cases of turbulent boundary layers with moderate pressure gradient (\cite{Klebanoff_1}, \cite{Klebanoff_2}, \cite{Smith}). This law, usually expressed as
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle U^+ = y^+, \ \ \ y^+ \le 5,
\end{array}
\label{wall law 00}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle U^+ \simeq A \ln y^+ +B, \ \ \ \ \ 5 < y^+ < 30,
\end{array}
\label{wall law 0}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle U^+ = \frac{1}{k} \ln y^+ + C, \ \ \ \ y^+ > 30,
\end{array}
\label{wall law}
\end{eqnarray}
gives the average velocity in different regions near the wall,
where $y^+=y U_T/\nu$ and $U^+ = U/U_T$ are the dimensionless normal coordinate and average velocity, and $U_T = \sqrt{\nu (\partial U/\partial y})_0$ is the friction velocity.
These expressions and the values of $A$, $B$, $C$ and $k$, seem to be universal properties of the flow, which do not depend on the Reynolds number. With reference to Fig. \ref{figura_1}, these equations, obtained through considerations of dimensional analysis and of self-similarity, hold under the hypothesis of fully developed parallel flow along the streamwise direction $x$ (\cite{Karman30}).
Specifically, Eq. (\ref{wall law 00}), being the direct consequence of the wall boundary condition and of the definition of $U_T$, expresses the velocity distribution in the laminar sub-layer LL, a domain adjacent to the wall where the effects of the viscosity are dominant.
Equation (\ref{wall law}) holds in the turbulent region TR, a zone of non-homogeneous turbulence, where the Taylor scale Reynolds number is high and variable with $y$.
Into Eq. (\ref{wall law}), $k$ is the von K\'arm\'an constant which according to
\cite{Karman30} and in line with Eq. (\ref{wall law}), can also be expressed as
\begin{eqnarray}
\displaystyle k = - \lim_{y^+ \rightarrow y^+_e}
\frac{\displaystyle \frac{d^2 U^+}{d {y^+}^2}}{\displaystyle \left( \frac{ d U^+}{d y^+} \right)^2}
\label{vk alt}
\end{eqnarray}
where $y^+_e$ defines the lower limit of the turbulent region.
As far as Eq. (\ref{wall law 0}) is concerned, this describes the velocity variations in the buffer layer BL, an intermediate zone between TR and LL in which viscous and inertia forces are comparable as order of magnitude.
$k$ and $A$ are directly related to the velocity variations along $y$, whereas $C$ and $B$ express the order of magnitude of $U$ with respect to the friction velocity in case of smooth wall.
These constants are dimensionless free parameters which can not be theoretically calculated (\cite{Landau}), therefore several experiments dealing with their determination were carried out (\cite{Fernholz}, \cite{Zagarola}), and scaling law similarities were proposed to justify Eq. (\ref{wall law}) (\cite{Barenblatt}).
These constants can be also identified through the elaboration of the results
of direct numerical simulations of the Navier-Stokes equations (\cite{Spalart},
\cite{Fernholz} and references therein).
From the various sources of the literature,
$k = 0.36 \div 0.44$,
$C = 4.5 \div 7.5$,
$A \simeq 5 \div 6$ and
$B \simeq -5 \div -3$.
Another law of the wall, also valid under the conditions of parallel flow and self-similarity, and that seems to show properties of universality, pertains the Reynolds stress $\langle u_x u_y \rangle$
\begin{eqnarray}
- \langle u v \rangle^+ = g(y^+), \ \ \
\langle u v \rangle^+ = \langle u_x u_y \rangle /U_T^2
\label{uv classic}
\end{eqnarray}
This behaves like $g \approx {y^+}^3$ very near the wall, and
at about $y^+ = 5$, exhibits an inflection point whose coordinate provides the
order of magnitude of the laminar sub-layer tickness.
The value $-\langle u v \rangle^+$ = 0.5, achieved in the buffer layer at about
$y^+ = 10 \div 13$, corresponds to the maximum turbulent energy production due to the mean flow, whereas far from the wall $g \approx 1$ (\cite{Hinze}, \cite{Tennekes}).
\begin{figure}[t]
\centering
\includegraphics[width=0.60\textwidth]{scheme_bl}
\caption{Schematic of the boundary layer.}
\label{figura_1}
\end{figure}
Although all these distributions seem to be universal laws,
according to \cite{Landau} and \cite{Fernholz}, there is not an adequate theory based on physical conjectures, other than the direct numerical simulations of the Navier Stokes equations, which leads to the calculation of the free parameters of these laws.
This represents the main motivation of the present work, whose purpose is to apply
the finite scale Lyapunov theory, described in \cite{deDivitiis_3}, to determine the wall velocity distribution and its parameters.
\bigskip
The first part of the work summarizes the main results of \cite{deDivitiis_3}, which deals with the steady homogeneous turbulence with uniform velocity gradient, and then shows that such results can be extended also to the non-homogeneous turbulence with shear rate.
We give a reasonable demonstration that the equation of the correlation function obtained by \cite{deDivitiis_3} can hold also in the case of non-homogeneous turbulence, in particular in the turbulent zone of the boundary layer.
This result allows to express, in the turbulent region, the several quantities such as
dimensionless velocity gradient and turbulent kinetic energy
in function of the local Taylor-scale Reynolds number, whereas
in the laminar sub-layer and in the buffer region, adequate variation laws of the
variables are assumed, which satisfy the wall boundary conditions
and match the values of the variables in the adjacent domains.
The analysis gives the von K\'arm\'an constant in terms of the variables at
the edge of the turbulent region, and provides a logarithmic velocity profile in TR,
different from Eq. (\ref{wall law}), which exhibits free parameters ($k$ included).
The knowledge of the statistical properties of the spanwise correlation function leads to the estimation of the variables at $y^+_e$ and thus to identify $k$ and the other parameters.
Therefore, the von K\'arm\'an is here theoretically obtained, resulting to be about 0.4,
in line with experiments and numerical simulations, and the other results, in agreement with the several data from the literature, show that the finite scale Lyapunov theory can be an adequate tool for studying the wall turbulence.
\bigskip
\section{\bf Resume }
\label{Resume}
This section summarizes the main results of \cite{deDivitiis_3} which regards the steady homogeneous turbulence in the presence of an average velocity gradient $\nabla_{\bf x} \bf U$.
There, the author, applying the Finite-scale Lyapunov theory (\cite{deDivitiis_1}) and the Liouville theorem, proposes the following evolution equation of the pair distribution function $F^{(2)}$ of fluid velocities in case of arbitrary flow
\begin{eqnarray}
\displaystyle \frac{\partial F^{(2)} }{\partial t}
+ \nabla_{{\bf x}} F^{(2)} \cdot {\bf v}
+ \nabla_{{\bf x}'} F^{(2)} \cdot {\bf v}'
= \lambda(r) \left( F^{(2)}_0 - F^{(2)} \right) -J_D
\label{3}
\end{eqnarray}
where $F^{(2)}_0$ is the pair distribution function of the isotropic turbulence which exhibits the same momentum and kinetic energy of $F^{(2)}$,
whereas $-J_D$ represents the rate of $F^{(2)}$ caused by the rate of the turbulent kinetic energy.
$\lambda (r)$ is the finite-scale Lyapunov exponent associated to the finite-scale
$r = \vert {\bf x}'-{\bf x} \vert$,
and $\bf v$ and $\bf v'$ are the fluid velocities calculated at $\bf x$ and $\bf x'$ respectively.
In case of homogeneous turbulence, \cite{deDivitiis_3} shows that the steady distribution function reasonably tends to a quantity which depends upon $F^{(2)}_0$, $\nabla_{\bf x} {\bf U}$ and $\lambda$
\begin{eqnarray}
\displaystyle F^{(2)} = F^{(2)}_0 + \frac{1}{\lambda(r)}
\left(
\frac{\partial F^{(2)}_0}{\partial v_j} \frac{\partial U_j}{\partial x_p} v_p
+ \frac{\partial F^{(2)}_0}{\partial v'_j} \frac{\partial U_j}{\partial x_p} v'_p
\right)
\label{F2}
\end{eqnarray}
The velocity correlation tensor $R_{k i} =\langle u_i u_j' \rangle$ is then calculated, by definition
\begin{eqnarray}
\begin{array}{c@{\hspace{+0.2cm}}l}
\displaystyle R_{k i} = \int_v \int_{v'} F^{(2)} u_k u'_i \ d^3u \ d^3u' =
R_{k i 0} - \frac{1}{\lambda}
\left( \frac{\partial U_k}{\partial x_p} R_{p i 0}
+ \frac{\partial U_i}{\partial x_q} R_{k q 0}
\right)
\end{array}
\label{R_0}
\label{R}
\end{eqnarray}
where $R_{k i 0}$ is the second order velocity correlation tensor associated to
the isotropic turbulence (\cite{Karman38}, \cite{Batchelor53})
\begin{eqnarray}
R_{k i 0} ({\bf r}) = u^2 \left( (f -g) \frac{r_k r_i}{r^2} + g \delta_{k i} \right)
\label{R0}
\end{eqnarray}
being $f$ and $g = f + 1/2 \ r \ \partial f/\partial r$ longitudinal and lateral
velocity correlation functions, respectively, and ${\bf u} = {\bf v} -{\bf U}$ is the
fluctuating velocity.
For $r = 0$, Eq. (\ref{R}) provides the expression of the Reynolds stresses
in function of $\nabla_{\bf x} {\bf U}$
\begin{eqnarray}
\left\langle u_k u_i \right\rangle = u^2 \left( \delta_{k i}
-\frac{1}{\Lambda} \left( \frac{\partial U_k}{\partial x_i}
+ \frac{\partial U_i}{\partial x_k} \right) \right)
\label{Boussinesq}
\label{u_x}
\end{eqnarray}
Equation (\ref{Boussinesq}) is a Boussinesq closure of the Reynolds stress, being
$u = \langle u_i u_i\rangle /3$, $\left\langle . \right\rangle$ indicates the average calculated on the ensamble of the fluid velocity, and $\Lambda= \lambda(0)$ is the maximal Lyapunov exponent.
The Lyapunov theory presented in \cite{deDivitiis_1} shows that
$\Lambda = u/\lambda_T$, being $\lambda_T$ the Taylor scale.
Next, the condition of steady flow leads to the following ordinary differential equation
for $f$ \cite{deDivitiis_3}
\begin{eqnarray}
\begin{array}{<EMAIL>}}l}
\displaystyle \sqrt{\frac{1-f}{2}} \ \frac{d f} {d \hat{r}} +
\displaystyle \frac{2}{R_T} \left( \frac{d^2 f} {d \hat{r}^2} +
\displaystyle \frac{4}{\hat{r}} \frac{d f}{d \hat{r}} \right) + \frac{10}{R_T}
\frac{f \ \hat{r}}{\sqrt{2(1-f)}} = 0
\end{array}
\label{I2 steady}
\end{eqnarray}
whose boundary conditions can be reduced to the following conditions of $f$
in the origin $r=0$
\begin{eqnarray}
\displaystyle f(0)=1, \ \ \ \frac{d f (0)} {d \hat{r}} = 0
\label{bc3}
\end{eqnarray}
where $\hat{r} = r/\lambda_T$, $d^2f/d\hat{r}^2 (0)\equiv -1$,
and $R_T = u \lambda_T /\nu $ is the Taylor scale Reynolds number.
Thus, Eqs. (\ref{I2 steady}) and (\ref{bc3}) express an initial condition problem
whose initial condition is given by Eqs. (\ref{bc3}).
The same steady condition gives the relationship between
$\nabla_{\bf x} \bf U$, $\Lambda$ and $R_T$, which represents the equation of the kinetic
energy in the case of steady homogeneous turbulence
\begin{eqnarray}
\frac{S} {\Lambda^2} = \frac{15}{R_T}
\label{steady E}
\end{eqnarray}
where $S$ is related to $\nabla_{\bf x} \bf U$
\begin{eqnarray}
S = \frac{\partial U_i}{\partial x_k}
\left( \frac{\partial U_k}{\partial x_i}
+ \frac{\partial U_i}{\partial x_k} \right)
\end{eqnarray}
The tensor $R_{i k}$ is then calculated with Eq. (\ref{R}) and (\ref{R0}), by means of $f$.
\cite{deDivitiis_3} studies some of the properties of Eqs. (\ref{I2 steady})-(\ref{bc3}),
and shows that their solutions behave like $f -1 \approx r$ in a given interval of $r$. Accordingly, the statistical moments of velocity difference are
$\langle \Delta u^n \rangle \approx r^{n/2}$ for $n <5$,
and the energy spectrum $E(\kappa) \approx k^{-2}$ in the inertial subrange.
Equation (\ref{3}) was obtained for an arbitrary flow,
whereas the other equations were derived under the assumption
of steady homogeneous turbulence with a given average velocity gradient.
The successive sections show that these equations can be applied also in the case of non-homogeneous turbulence where $\lambda_T$ and $u$ vary with the space coordinates.
Thus, the method is here applied to the turbulent region of the fully developed boundary layer, where the several quantities vary with the wall normal coordinate.
\bigskip
\section{\bf Analysis \label{Analysis}}
An evolution equation for the velocity correlation
is determined, in the case of non-homogeneous turbulence with a nonzero
average velocity gradient.
The fluid velocity, measured in the reference frame $\Re$, is
${\bf v} ={\bf U} + {\bf u}$, where ${\bf U} \equiv (U_x, U_y, U_z)$ and
${\bf u} \equiv (u_x, u_y, u_z)$ are, average and fluctuating velocity, respectively.
The velocity correlation tensor is $R_{i j} = \langle u_i u'_j \rangle$, where
$u_i$ and $u'_j$ are the velocity components of $\bf u$ calculated at $\bf x$
and ${\bf x'} = {\bf x} + {\bf r}$, and $\bf r$ is the separation distance.
As the analysis is finalized to the description of the turbulent region of the developed boundary layer, and since there the effects of the spatial variations of
$\nabla_{\bf x} {\bf U}$ are orders of magnitude much smaller than those caused by
$\nabla_{\bf x} {\bf U}$ (\cite{Karman30}, \cite{Nikuradse}), the velocity gradient is assumed to be a function of $\bf x$ alone, being $\nabla_{\bf x} {\bf U} = \nabla_{\bf x} {\bf U}'$.
In order to determine the evolution equation of $R_{i j}$, the Navier-Stokes equations are written for the fluctuating velocity in the points ${\bf x}$ and ${\bf x}'$.
The evolution equation of $R_{i j}$ is determined by multiplying first and second equation by $u'_j$ and $u_i$, respectively, summing the so obtained equations, and calculating the average on the statistical ensemble (\cite{Karman38}, \cite{Batchelor53})
\begin{eqnarray}
\displaystyle
\frac{\partial R_{i j}}{\partial t}=
T_{i j} + P_{i j}
+ 2 \nu \nabla^2 R_{i j}
- \frac{\partial U_i}{\partial x_k} R_{k j}
- \frac{\partial U_j}{\partial x_k} R_{i k}
+ \frac{\partial R_{i j} }{\partial r_k}
(U_k - U_k') + \Gamma_{i j}
\label{cc1 A}
\label{cc1}
\end{eqnarray}
being
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle T_{i j} ({\bf x}, {\bf r}) =\frac{\partial }{\partial r_k}
\left\langle u_i u_j' (u_k - u_k') \right\rangle, \ \ \ \
\displaystyle P_{i j} ({\bf x}, {\bf r}) = \frac{1}{\rho}
\frac{\partial \langle p u'_j \rangle}{\partial r_i}
\end{array}
\end{eqnarray}
and $p$ is the fluctuating pressure.
The quantities of Eq. (\ref{cc1}), which in turn depend on $\bf r$, due to non-homogeneity depend also on $\bf x$, and Eq. (\ref{cc1}) incudes an additional term with respect to the homogeneous turbulence, represented by $\Gamma_{i j}({\bf x}, {\bf r})$ (\cite{Oberlack})
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \Gamma_{i j} ({\bf x}, {\bf r})=
-\frac{1}{\rho} \frac{\partial \langle p u_j' \rangle }{\partial x_i}
-\frac{\partial}{\partial x_k } \langle u_i u_k u_j' \rangle
- \frac{1}{\rho} \frac{\partial \langle p' u_i \rangle }{\partial r_j}
\displaystyle + \nu \frac{\partial^2 R_{i j} }{\partial x_k \partial x_k } \\\\
\displaystyle -2 \nu \frac{\partial^2 R_{i j} }{\partial r_k \partial x_k }
\end{array}
\end{eqnarray}
which provides the non-homogeneity of the different terms of correlation.
Making the trace of Eq. (\ref{cc1 A}), we obtain the following scalar equation
\begin{eqnarray}
\displaystyle
\frac{\partial R}{\partial t}=
\frac{1}{2} H
+ 2 \nu \nabla^2 R
- \frac{\partial U_i}{\partial x_k} R_{i k}^S
+ \frac{\partial R }{\partial r_k}
(U_k - U_k') + \Gamma
\label{cc2}
\end{eqnarray}
being $R_{i k}^S$ is the symmetric part of $R_{i k}$ and
\begin{eqnarray}
\displaystyle R = \frac{R_{i i}}{2} , \ \ \ \Gamma = \frac{\Gamma_{i i}}{2}
\end{eqnarray}
$R({\bf x}, 0)$ gives the turbulent kinetic energy, $H$ $\equiv$ $T_{i i}$ provides the mechanism of energy cascade (\cite{Karman38}, \cite{Batchelor53}),
$P_{i i} \equiv 0$ expresses the fluid incompressibility whereas $\Gamma$
arises from the non-homogeneity of the flow.
Now, a scalar equation for describing the
main properties of the velocity correlation is determined.
To this end, $R_{i j}$, $H$ and $\bf U$ are decomposed into an even function of
$r \equiv \vert {\bf r} \vert$ (here called spherical part), plus the remaining term:
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
R_{i j} ({\bf x}, {\bf r})= \hat{R}_{i j}({\bf x}, r) + \Delta R_{i j} ({\bf x}, {\bf r}) \\\\
H ({\bf x}, {\bf r}) = \hat{H}({\bf x}, r) + \Delta H ({\bf x}, {\bf r}) \\\\
{\bf U}' -{\bf U} = \hat{\bf U}({\bf x}, r) + \Delta {\bf U} ({\bf x}, {\bf r})
\end{array}
\label{dec}
\end{eqnarray}
where $\hat{F}$ is the spherical part of the generic quantity $F$, defined as
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.99cm}}l}
\displaystyle \hat{F}({\bf x}, r) = \frac{1}{6} \left( F({\bf x}, r, 0, 0) + F({\bf x}, 0, r, 0)+ F({\bf x}, 0, 0, r) \right) \\\\
\displaystyle + \frac{1}{6} \left( F({\bf x}, -r, 0, 0) + F({\bf x}, 0, -r, 0)+ F({\bf x}, 0, 0, -r) \right)
\end{array}
\label{isotrp}
\end{eqnarray}
and $\Delta R_{i j}({\bf x}, {\bf 0})$=$\Delta H({\bf x}, {\bf 0})$ = 0.
Therefore, the Fourier transform of $\hat{R}$ identifies the part of the energy spectrum depending upon $\mbox{\boldmath $\kappa$}^2$.
As the consequence of this decomposition, $R$ and $\Delta R$ satisfy the equations
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{\partial \hat{R}}{\partial t} =
\frac{\hat{H}}{2}
+ 2 \nu \left( \frac{\partial^2 \hat{R}} {\partial r^2} +
\displaystyle \frac{2}{r} \frac{\partial \hat{R}}{\partial r} \right)
- \hat{G}
\end{array}
\label{eq iso}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{\partial \Delta R}{\partial t} =
\frac{ \Delta H}{2}
+ 2 \nu \nabla^2 \Delta R
- \Delta{G}
\end{array}
\label{eq non-iso}
\end{eqnarray}
in which Eq. (\ref{eq non-iso}) is obtained as the difference between Eqs.
(\ref{cc2}) and (\ref{eq iso}), and
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \hat{G} = \frac{\partial U_i} {\partial x_k} \hat{R}_{k i}^S + \hat{G}_0, \ \ \ \
\displaystyle \Delta{G} = \frac{\partial U_i} {\partial x_k} \Delta{R^S}_{k i}
- \Delta \left( \frac{\partial R }{\partial r_k} (U_k - U_k')\right) -\Delta \Gamma
\end{array}
\end{eqnarray}
and $\hat{G}_0$ represents the spherical part of $-\partial R /\partial r_k (U_k - U_k') -\Gamma $.
It is worth to remark that Eq. (\ref{eq iso}) formally coincides with the equation obtained
in \cite{deDivitiis_3},
with the difference that here, because of non-homogeneity, the quantities appearing into
Eq. (\ref{eq iso}) depend also on $\bf x$.
The turbulence is here studied using Eq. (\ref{eq iso})
alone, whereas $R_{i j}({\bf x}, {\bf r})$ will be determined in function of
$\nabla_{\bf x}{\bf U}$, by means of a proper statistical analysis
of the two-points velocity correlation.
\bigskip
\section{\bf Pair distribution function}
\label{distribution function}
This section analyses the non-homogeneous turbulence through the pair distribution function, taking into account that $\lambda_T$ and $\langle u_i u_j \rangle$ vary with the spatial coordinates, whereas $\nabla_{\bf x} {\bf U}$ is considered to be an assigned quantity.
To study this, consider now the pair distribution function of the fluid velocity
\begin{eqnarray}
F^{(2)} ({\bf v}, {\bf v}'; {\bf x}, {\bf x}') = F^{(2)}_0 \left( {\bf v}, {\bf v}'; {\bf x}, {\bf x}' \right) + \phi^{(2)} ({\bf v}, {\bf v}'; {\bf x}, {\bf x}')
\label{iso00}
\end{eqnarray}
where $F^{(2)}$ obeys to Eq. (\ref{3}), and $\phi^{(2)}$, representing the deviation from the isotropic turbulence, satisfies, at each instant, the following equations
(\cite{deDivitiis_3})
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \int_{v} \int_{v'} \phi^{(2)} du^3 du'^3= 0, \\\\
\displaystyle \int_{v} \int_{v'} \phi^{(2)} {\bf u} \ du^3 du'^3= 0, \\\\
\displaystyle \int_{v} \int_{v'} \phi^{(2)} {\bf u} \cdot {\bf u} \ du^3 du'^3= 0.
\end{array}
\label{cons1}
\end{eqnarray}
Equations (\ref{cons1}) state that momentum and kinetic energy associated to $F^{(2)}_0$ and $F^{(2)}$ are equal each other, respectively.
The analysis supposes also that all the dimensionless statistical moments of $F^{(2)}_0$ are constant in all the points of the fluid domain, therefore
the functional form of $F^{(2)}_0$ is assumed to be
\begin{eqnarray}
F^{(2)}_0 ({\bf v}, {\bf v}'; {\bf x}, {\bf x}') =
F^{(2)}_0 \left( \frac{{\bf v} -{\bf U} ({\bf x})}{u({\bf x})}, \frac{{\bf v}'-{\bf U} ({\bf x}')}{u( {\bf x}')} \right)
\label{iso0}
\end{eqnarray}
and its gradients are calculated in functions of the spatial derivatives of $\bf U$ and $u$
\begin{eqnarray}
\frac{\partial F^{(2)}_0}{\partial x_k} =
- \frac{\partial F^{(2)}_0}{\partial v_j}
\left( \frac{\partial U_j}{\partial x_k} + \frac{u_j}{u} \frac{\partial u}{\partial x_k}\right) ,
\ \ \ \
\frac{\partial F^{(2)}_0}{\partial x'_k} =
-\frac{\partial F^{(2)}_0}{\partial v'_j}
\left( \frac{\partial U_j}{\partial x_k} + \frac{u_j'}{u'} \frac{\partial u'}{\partial x_k'}\right)
\label{4}
\end{eqnarray}
where, as before, $\nabla_{\bf x} {\bf U} = \nabla_{\bf x'} {\bf U}'$.
Substituting Eqs.(\ref{iso00}) and (\ref{4}) into Eq. (\ref{3}), we obtain the following relationship between $F^{(2)}_0$ and $\phi^{(2)}$
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \lambda \phi^{(2)} =
-J_D
\displaystyle - \left( \frac{\partial F^{(2)}_0 }{\partial t}
+ \frac{\partial \phi^{(2)} }{\partial t}
+ \frac{\partial \phi^{(2)} }{\partial x_p} v_p
+ \frac{\partial \phi^{(2)} }{\partial x_p'} v_p'
\right) \\\\
\hspace{+15 mm}
\displaystyle +\left(
\frac{\partial F^{(2)}_0}{\partial v_j} \frac{v_p u_j}{u} \frac{\partial u}{\partial x_k}\right)
+ \left(\frac{\partial F^{(2)}_0}{\partial v'_j} \frac{ v'_p u'_j}{u'} \frac{\partial u'}{\partial x'_k} \right)
\\\\
\hspace{+15 mm}
\displaystyle +\left(
\frac{\partial F^{(2)}_0}{\partial v_j} v_p
+ \frac{\partial F^{(2)}_0}{\partial v'_j} v'_p
\right) \frac{\partial U_j}{\partial x_p}
\end{array}
\label{fi_2}
\end{eqnarray}
As the last term of Eq. (\ref{fi_2})
identically satisfies Eqs. (\ref{cons1}), these latter are written in the form
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \int_v \hspace{-1.0mm} \int_{v'} \hspace{-2.0mm}
\left[\begin{array}{c}
\displaystyle 1 \\\
\displaystyle {\bf u} \\\
\displaystyle {\bf u} \cdot {\bf u}
\end{array}\right]
\displaystyle ( \frac{\partial F^{(2)} }{\partial t}
+ \frac{\partial \phi^{(2)} }{\partial x_p} v_p
+ \frac{\partial \phi^{(2)} }{\partial x_p'} v_p'
- \frac{\partial F^{(2)}_0}{\partial v_j} \frac{v_p u_j}{u}
\frac{\partial u}{\partial x_k} \\\\
\displaystyle - \frac{\partial F^{(2)}_0}{\partial v'_j} \frac{ v'_p u'_j}{u'} \frac{\partial u'}{\partial x'_k}
+ J_D ) \ du^3 du'^3 \equiv 0
\end{array}
\label{c0}
\end{eqnarray}
A sufficient condition for satisfying these five equations is that the integrand of
Eq. (\ref{c0}) vanishes
\begin{eqnarray}
\frac{\partial F^{(2)} }{\partial t}
+ \frac{\partial \phi^{(2)} }{\partial x_p} v_p
+ \frac{\partial \phi^{(2)} }{\partial x_p'} v_p'
- \frac{\partial F^{(2)}_0}{\partial v_j} \frac{v_p u_j}{u} \frac{\partial u}{\partial x_k}
- \frac{\partial F^{(2)}_0}{\partial v'_j} \frac{ v'_p u'_j}{u'} \frac{\partial u'}{\partial x'_k}
+ J_D \equiv 0
\label{sc}
\end{eqnarray}
Assuming that Eq. (\ref{sc}) is true, and taking into account Eq. (\ref{fi_2}),
$F^{(2)}$ is given by Eq. (\ref{F2}) which does not depend on the gradient of $u$ and $u'$,
as in the case of the homogeneous turbulence.
Really, $F^{(2)}$ changes starting from an arbitrary initial condition, therefore
Eq. (\ref{F2}) represents an approximation which can be considered to be valid far from the initial condition.
As the consequence, also the spherical part of the correlation tensor, $\hat{R}_{i j}$, here obtained
\begin{eqnarray}
\hat{R}_{k i} = \hat{R}^S_{i k} = \frac{u^2}{3} \left(3 f + \frac{\partial f}{\partial r} r \right) \left( \delta_{k i}
-\frac{1}{\lambda} \left( \frac{\partial U_k}{\partial x_i}
+ \frac{\partial U_i}{\partial x_k} \right) \right)
\label{R2}
\end{eqnarray}
coincides with the expression given in \cite{deDivitiis_3}.
Thus, substituting Eq. (\ref{R2}) and into Eq. (\ref{eq iso}) and following the analytical procedure of \cite{deDivitiis_3}, we found that $f$ obeys to Eq. (\ref{I2 steady}) also in the present case, and $\Lambda$ is related to $\nabla_{\bf x} {\bf U}$ and $R_T$ through Eq. (\ref{steady E}). The difference with respect to \cite{deDivitiis_3} is that, here the turbulence is non-homogeneous, thus $R_T = R_T({\bf x})$, and $f$ also depends on $\bf x$, being
$f = f({\bf x}, r)$.
\bigskip
\section{\bf Analysis of the boundary layer}
Here, the steady turbulent boundary layer with a moderate pressure gradient is analyzed,
assuming that the flow is fully developed along the streamwise direction.
To this purpose, return to Fig. \ref{figura_1},
and consider only the developed region of the flow.
In the figure, $\Re$ is the wall frame of reference, $x$ and $y$ are, respectively, the streamwise direction and the coordinate normal to the wall, whereas $z$ is the spanwise coordinate.
In the developed region, $u$, $\langle u_x u_y \rangle$ and $\lambda_T$
change with $y$, and in the laminar sublayer $U_x$ and $u$ are both about proportional to $y$.
In LL, BL and at the beginning of TR, the correlation scale of velocity is proportional to the distance from the wall, and vanishes for $y=0$, being
\begin{eqnarray}
\displaystyle \lambda_T = \left( \frac{\partial \lambda_T}{\partial y} \right)_0 y + ...
\label{lambda e}
\end{eqnarray}
where $\left( {\partial \lambda_T}/{\partial y} \right)_0 = O(1)$, whereas
the velocity fluctuations follow the Navier-Stokes equations and satisfy the wall
boundary conditions
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle u_x = \frac{\partial u_x}{\partial y}(0) y + ..., \ \ \
\displaystyle u_y = \frac{1}{2} \frac{\partial^2 u_y}{\partial y^2}(0) y^2 + ..., \ \ \
\displaystyle u_z = \frac{\partial u_z}{\partial y}(0) y + ...
\end{array}
\label{bc}
\end{eqnarray}
In case of fully developed flow along $x$, that is, parallel flow assumption
($\partial / \partial y$ $>>>$ $\partial / \partial x$), the continuity and momentum equations
of the mean flow are (\cite{Schlichting})
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{\partial U_y}{\partial y} =0,
\end{array}
\label{continuity 2}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{\partial}{\partial y} \langle u_x u_y \rangle
\displaystyle + \frac{1}{\rho} \frac{\partial P}{\partial x}
\displaystyle - \nu \frac{\partial^2 U_x}{\partial y^2}=0, \\\\
\displaystyle \frac{\partial}{\partial y} \langle u_y^2 \rangle
\displaystyle + \frac{1}{\rho} \frac{\partial P}{\partial y} =0
\end{array}
\label{momentum 2}
\end{eqnarray}
where $P$ is the average pressure, whereas the equation of the turbulent kinetic energy reads as
\begin{eqnarray}
\displaystyle \frac{\partial}{\partial y} \left\langle u_y \left( \frac{p}{\rho}
+ \frac{u_j u_j}{2}\right) \right\rangle
\displaystyle + \left\langle u_x u_y \right\rangle \frac{\partial U_x}{\partial y}
\displaystyle + \nu \left\langle \frac{\partial u_j}{\partial x_i} \frac{\partial u_j}{\partial x_i} \right\rangle =0
\label{energy 2}
\end{eqnarray}
The non-homogeneity is responsible for the first term
of Eq. (\ref{energy 2}), whereas second and third terms represent, respectively,
the energy production due to the average motion, and the dissipation.
\bigskip
Because of the parallel flow assumption, in all these equations $U_x$, $\langle u_i u_j\rangle$ are functions of $y$ alone.
From Eq. (\ref{continuity 2}) and taking into account the boundary conditions, $U_y \equiv 0$,
whereas Eqs. (\ref{momentum 2}) give the following first integrals
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{P(x, y)}{\rho} = F x + H - \langle u_y^2 \rangle
\end{array}
\label{y}
\end{eqnarray}
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \langle u_x u_y \rangle -\nu \frac{d U_x}{d y}=
- F y -U_T^2
\end{array}
\label{x}
\end{eqnarray}
being
$F= 1/\rho \ \partial P/ \partial x$
and $H$ is a proper constant proportional to the average pressure at $x=0$.
Into Eq. (\ref{x}), the boundary layer approximation and the hypothesis
of moderate average pressure gradient along $x$ provide that
$ U_T^2 >> \vert F y \vert$ in all the regions, therefore introducing the dimensionless variables $U^+ = U_x/U_T$ and $y^+ = y U_T /\nu$, Eq. (\ref{x}) reads as
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \frac{d U^+}{d y^+}= \frac{1}{1+R_T}
\end{array}
\label{x1}
\end{eqnarray}
Equation (\ref{x1}) is assumed to describe the average flow with
moderate pressure gradient in the three regions of the boundary layers,
when the parallel flow hypothesis is verified.
\bigskip
\subsection{\bf The Turbulent Region}
This section studies the distribution of the different variables in
the turbulent region TR, by means of the analysis seen in the sections \ref{Analysis} and \ref{distribution function}.
First, observe that, in case of fully developed parallel flow,
the effects of non-homogeneity in TR are much smaller than those related
to energy production and dissipation (\cite{Tennekes}), thus the first term of Eq. (\ref{energy 2}) is
here neglected with respect to the other ones, and Eq. (\ref{steady E}) is recovered.
This approximation, in agreement with the analysis seen in sect. \ref{distribution function}, states that the kinetic energy production is balanced only by the dissipation in TR, and allows to express the several variables in terms
of the local value of $R_T$ (or $dU_x/dy$).
In order to calculate $u^+ = u/U_T$, $dU_x/dy$ is eliminated between
Eqs. (\ref{x}) and (\ref{steady E}), where $S = (dU_x/dy)^2$
\begin{eqnarray}
\displaystyle u^+ = \frac{R_T^{3/4}}{15^{1/4} \sqrt{1+R_T}}
\label{u}
\end{eqnarray}
Being $R_T = u^+ \lambda_T^+$, also $\lambda_T^+$ is in terms of $R_T$
\begin{eqnarray}
\displaystyle \lambda_T^+ = (15 R_T)^{1/4} \sqrt{1+R_T}
\label{lambda}
\end{eqnarray}
As these equations arise from Eq. (\ref{steady E}) which holds in the turbulent
region, Eqs.(\ref{u}) and (\ref{lambda}) describe the variations of $u$ and $\lambda_T$
only in TR.
From the comparison between Eqs. (\ref{x1}) and (\ref{vk alt}), $\left( {d R_T}/{d y^+}\right)_e$ identifies the von K\'arm\'an constant, here expressed taking into account that
$R_T \equiv \lambda_T^+ u^+$
\begin{eqnarray}
k \equiv \left( \frac{d R_T}{d y^+}\right)_e =
\left( \frac{d u^+}{d y^+}\right)_e \lambda_{T e}^+ +
\left( \frac{d \lambda_{T}^+}{d y^+}\right)_e u_e^+
\label{vkc}
\end{eqnarray}
where the subscript $e$ denotes the values calculated at the edge of TR.
Substituting Eqs. (\ref{u}) and (\ref{lambda}) into Eq. (\ref{vkc}), we obtain the von K\'arm\'an constant in function of the variables at $y^+_e$
\begin{eqnarray}
\displaystyle k = \frac{4}{15^{1/4}} \left( \frac{d \lambda_{T}^+}{d y^+}\right)_e
\displaystyle \frac{\sqrt{1+R_{T e}}}{1+ 3 R_{T e}} R_{T e}^{3/4}
\label{vkc1}
\end{eqnarray}
It is worth to remark that this estimation of $k$ requires the knowledge of $R_{T e}$ and
of $(d\lambda_T^+/dy^+)_e$, whereas does not
need the assumption that $U^+$ is represented by a logarithmic profile.
Next, the dimensionless Prandtl's mixing scale $l_p^+$ is calculated from the definition
of Prandtl's mixing length $l_p$
\begin{eqnarray}
- \langle u_x u_y \rangle = \left| \frac{\partial U_x}{\partial y} \right|
\frac{\partial U_x}{\partial y} \ l_p^2
\label{Prandtl_d}
\end{eqnarray}
and Eqs. (\ref{u_x}) and (\ref{steady E}). This is
\begin{eqnarray}
l_p^+ = \sqrt{R_T (1 + R_T)}
\label{Prandtl}
\end{eqnarray}
being $l_p^+ \simeq R_T$ for $R_T >> 1$, and its derivative calculated at $y^+_e$ is also
expressed in terms of $R_{T e}$
\begin{eqnarray}
\displaystyle \left( \frac{d l^+_p}{d y^+} \right)_e = \frac{1+2 R_{T e}}{2 \sqrt{R_{T e}(1+R_{T e})}} \ k
\label{Prandtl_y}
\end{eqnarray}
In view of Eq. (\ref{vkc1}), this derivative can be also expressed in function of $d \lambda_T^+ /d y^ì$
\begin{eqnarray}
\displaystyle \left( \frac{d l^+_p}{d y^+} \right)_e = 2 \ \frac{1+2 R_{T e}}{1+ 3 R_{T e}} \
\left( \frac{ R_{T e} }{15}\right)^{1/4} \left( \frac{d \lambda_{T}^+}{d y^+}\right)_e
\displaystyle
\label{Prandtl_y1}
\end{eqnarray}
This expression gives the link between the variations of the Taylor scale and of the Prandtl's length at the border of the turbulent region.
Following Eqs. (\ref{u}) and (\ref{lambda}), $u^+$ and $\lambda_T^+$ are functions of $y^+$ through the local value of $R_T$ (or of $dU^+/dy^+$), therefore, the distribution of such quantities along $y^+$ require the knowledge of the function $R_T=R_T(y^+)$.
This latter can be expressed as
\begin{eqnarray}
\displaystyle R_T(y^+) = R_{T e} + \left( \frac{d R_T}{d y^+}\right)_e
\left( y^+ - y^+_e \right) + O \left( y^+ - y^+_e \right)^2
\label{R_T}
\end{eqnarray}
Now, in a range of $y^+$ where $O \left( y^+ - y^+_e \right)^2$ is negligible with respect to the other terms, $d U^+/d y^+$ is
\begin{eqnarray}
\displaystyle \frac{d U^+}{d y^+}= \displaystyle \frac{1}{\displaystyle 1 + R_{T e} + \left( \frac{d R_T}{d y^+}\right)_e
\displaystyle \left( y^+ - y^+_e \right)}
\label{vd}
\end{eqnarray}
$U^+$ exhibits there logarithmic law, obtained integrating Eq. (\ref{vd})
from $y^+_e$ to $y^+$
\begin{eqnarray}
U^+ = \frac{1}{k} \ln \left( \frac{1+R_{T e} + k (y^+-y^+_e)}{1+R_{T e}} \right) + U^+_e
\label{U turbo}
\end{eqnarray}
being $U^+_e=U^+(y^+_e)$.
This law is defined as soon as the parameters $y^+_e$, $k$ and $U^+_e$
are known. Equation (\ref{U turbo}) differs from the classical expression (\ref{wall law})
and formally tends to Eq. (\ref{wall law}) when $y^+ \rightarrow \infty$.
Therefore, Eq. (\ref{U turbo}) can give values of $U^+$ sizably different from (\ref{wall law}) for small $y^+$.
The comparison between these equations, for $y^+ \rightarrow \infty$ identifies
$C$ in terms of the variables at $y_e$
\begin{eqnarray}
C = U^+_e + \frac{1}{k} \ln \left( \frac{k}{1+R_{T e}}\right)
\label{C}
\end{eqnarray}
\bigskip
As far as the Reynolds stress is concerned, it is expressed in function of $R_T$
through Eqs. (\ref{x}) and (\ref{x1})
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle \langle u v \rangle^+ \equiv \frac{\langle u_x u_y \rangle}{U_T^2} = - \frac{R_T}{1+R_T}
\end{array}
\label{uv}
\end{eqnarray}
Observe that, the wall boundary conditions (\ref{bc})
state that $\langle u v \rangle^+ \approx {y^+}^3$ near the wall,
whereas Eq. (\ref{uv}) gives $\langle u v \rangle^+ \approx R_T \approx {y^+}^2$.
This disagreement is due to the fact that the expression of $\langle u v \rangle^+$
has not been derived from the correlation equation with $r=0$, but arises from
Eq. (\ref{x}) eliminating $dU_x/dy$. This implies that Eq. (\ref{uv}) holds only in TR and BL, whereas in the laminar sub-layer a proper matching condition must be applied.
\bigskip
\subsection{\bf Matching Turbulent region - buffer layer}
With reference to Fig. \ref{figura_1}, the domains LL and BL constitute SL,
a zone between wall and turbulent region.
There, due to the presence of the wall, the analysis of sections \ref{distribution function}
and \ref{Analysis} can not be applied.
Therefore, the mean variables in SL are expressed in function of $y^+$,
taking into account the boundary conditions (\ref{bc}) and that $\lambda_T$
follows Eq. (\ref{lambda e}), where it is assumed
$(d \lambda_T^+/dy^+)_e = (d \lambda_T^+/dy^+)_0$.
As the consequence, $R_T$, $u^+$
and $\lambda_T^+$ are supposed to vary in SL according to
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle R_T = u^+ \lambda_T^+, \\\\
\displaystyle u^+ = {u^+_e} \ \frac{y^+ +C_u {y^+}^2}{y^+_e +C_u {y^+_e}^2}, \ \ \ (0 < y^+ \le y^+_e) \\\\
\displaystyle \lambda_T^+ = \lambda_{T e}^+ \frac{y^+}{y^+_e}, \ \ \ (0 < y^+ \le y^+_e),
\end{array}
\label{ss}
\end{eqnarray}
where $C_u$ is a constant given by
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle C_u = \frac{1}{y^+_e} \frac{2 R_{T e}- k y^+_e}{k y^+_e -3 R_{T e}}
\end{array}
\label{ssc}
\end{eqnarray}
Equations (\ref{ss}) and (\ref{ssc}) provide the matching condition between SL and TR.
Specifically, Eqs. (\ref{ss}) state that $u^+$, $R_T$ and
$\lambda_T^+$ are continuous functions for $y^+ = y^+_e$, whereas
Eq. (\ref{ssc}) gives there the continuity of their derivatives.
The average velocity is then calculated by quadrature, substituting the expression of $R_T$ given by Eqs. (\ref{ss}), into Eq. (\ref{x1}),
integrating this latter from 0 to $y^+$, with $U^+(0) = 0$
\begin{eqnarray}
U^+(y^+) = \int_0^{y^+} \frac{1}{1+R_T(\eta^+)} d \eta^+
\label{U ss}
\end{eqnarray}
where $U^+(y^+) \simeq y^+$ for small $y^+$.
The matching between SL and TR gives the value of $U^+_e$.
\begin{eqnarray}
U^+_e = \int_0^{y^+_e} \frac{1}{1+R_T(\eta^+)} d \eta^+
\end{eqnarray}
Thus, Eq. (\ref{U ss}) and (\ref{U turbo}) establish that both $U^+$ and $d U^+ / d y^+$ are continuous
for $y^+ = y^+_e$
\bigskip
\subsection{\bf Matching buffer layer - laminar sub-layer}
As previously seen, the expression (\ref{uv}) of the Reynolds stress, valid
in BL and TR, can not be applied in the laminar region.
Since $\langle u v \rangle^+ \approx {y^+}^3$ near the wall,
the Reynolds stress in LL is approximated by
\begin{eqnarray}
\displaystyle \langle u v \rangle^+ = \langle u v \rangle^+_*
\displaystyle \frac{ {y^+}^3 +C_{uv} {y^+}^4 } {{y^+_*}^3 +C_{uv} {y^+_*}^4}, \ \ \ (0 < y^+< y^+_*)
\label{ss uv}
\end{eqnarray}
being $C_{uv}$ a constant
\begin{eqnarray}
\displaystyle C_{uv} = \frac{1}{y^+_*} \frac{3- (d \ln \langle uv \rangle^+/ d y^+)_*
y^+_*} {(d \ln \langle uv \rangle^+/ d y^+)_* y^+_* -4}
\label{ss uv c}
\end{eqnarray}
Equations (\ref{ss uv}) and (\ref{ss uv c}) establish that the Reynolds stress and
its derivative are continuous for $ y^+ = y^+_*$, where the subscript $*$ indicates
the value calculated at $ y^+_*$.
This latter, obtained as the inflection point of
$\langle u v \rangle^+$ (i.e. $d^2 \langle u v \rangle^+/{dy^+}^2 =0$) according to Eqs. (\ref{uv}) and (\ref{ss}), gives the dimensionless tickness of the laminar sub-layer.
\bigskip
\section{\bf Identification of the velocity law free parameters}
The definition of the velocity law, requires the knowledge of the
parameters which appear into Eqs. (\ref{vkc1}) and (\ref{U turbo}).
In particular, the determination of $k$ needs the values of $R_{T e}$ and
$(d \lambda_T^+/d y^+)_e$.
To identify these latter, we will proceed as follow.
First, observe that $\langle u_i u_j \rangle$ is a symmetric tensor which can be obtained from the diagonal tensor
\begin{eqnarray}
\langle {\bf u}_0 {\bf u}_0 \rangle =
\left[\begin{array}{ccc}
\displaystyle \langle u_0^2 \rangle & 0 & 0 \\\\
\displaystyle 0 & \langle v_0^2 \rangle & 0 \\\\
\displaystyle 0 & 0 & \langle w_0^2 \rangle
\end{array}\right]
\end{eqnarray}
through an opportune rotation around $y \equiv y_0$, being $\langle u_0^2 \rangle$, $\langle v_0^2 \rangle$ and $\langle w_0^2 \rangle$ the velocity components standard deviations in the canonical frame, and $\vartheta$ is the angle of this rotation.
Therefore, $u$ and $\langle u_x u_y \rangle$ are in terms of the elements of
$\langle {\bf u}_0 {\bf u}_0 \rangle$
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle u^2 = \frac{1}{3} (\langle u_0^2 \rangle + \langle v_0^2 \rangle +\langle w_0^2 \rangle) \\\\
\displaystyle \langle u_x u_y \rangle = (\langle u_0^2 \rangle - \langle v_0^2 \rangle) \frac{\sin 2 \vartheta}{2}
\end{array}
\label{cr}
\end{eqnarray}
Because of the parallel flow assumption and taking into account Eqs. (\ref{u_x}) and (\ref{cr}), $u$ and $\langle u_x u_y \rangle$ are related each other in such a way that
\begin{eqnarray}
u^2 \ge \vert \langle u_x u_y \rangle \vert
\end{eqnarray}
Hence, Eq. (\ref{steady E}) implies that
\begin{eqnarray}
R_T \ge 15 \ \ \ {\mbox{or}} \ \ \ \frac{d U_x}{d y} \lambda_T \ge u
\end{eqnarray}
in the turbulent region.
This limitation identifies the minimum value of $R_T$ in TR, which is assumed to be
\begin{eqnarray}
R_{T e} = 15.
\end{eqnarray}
To estimate $(d \lambda_T^+/d y^+)_e$, consider first the spanwise correlation function of the streamwise velocity components
(that is $R_{1 1} (r_z)$ = $\langle u_x(x, y, z) u_x(x, y, z+r_z) \rangle$).
This can be calculated with Eq. (\ref{R}), once $f$ is known through
Eqs. (\ref{I2 steady}).
Since $\partial U / \partial y$ leads to the development of coherent structures in the fluid, similar to streaks and caused by the stretching of the vortex lines along
the streamwise direction (\cite{Kim}), we expect that $R_{1 1} (r_z)$
intersects the horizontal axis and remains negative for $r_z \rightarrow \infty$.
This implies a wide distribution of spacings between the different streaky
structures, whose mean value depends on $R_T$ (\cite{Kim}).
Accordingly, there exists a minimum spanwise distance $r_{z}^*$
such that, for $r_{z} \in \left[0, r_{z}^* \right]$, $R_{1 1} (r_z)$ gives the necessary informations to describe the statistical properties of the correlation
(\cite{Ventsel}).
\begin{figure}[t]
\centering
\includegraphics[width=0.60\textwidth]{R11_scheme}
\caption{Schematic of the spanwise correlation function of $u_x$ at the edge of the turbulent region and definition of Eqs. (\ref{condizio 1}) and (\ref{condizio 2}).}
\label{figura_2}
\end{figure}
According to the theory (\cite{Ventsel}), if $R_{1 1}$ monotonically tends to zero as
$r_z \rightarrow \infty$ and $R_{1 1} < 0$, $r_{z}^*$ can be estimated as the distance at which $R_{1 1}$ is negative and exhibits an inflection point (second inflection point of the curve, see Fig. \ref{figura_2})
\begin{eqnarray}
\displaystyle R_{1 1}(r_{z}^*) < 0, \ \ \ \ \frac{\partial^2 R_{1 1}}{ \partial r_z^2} (r_{z}^*)=0,
\ \ \ \mbox{Condition 1}
\label{condizio 1}
\end{eqnarray}
Alternatively, $r_{z}^*$ can be estimated as the intersection between the osculating parabola in the point $\bar{r}_{z}$ where $R_{1 1}$ = min, and the horizontal axis, that is
\begin{eqnarray}
R_{1 1}(\bar{r}_{z}) +
\frac{1}{2} \frac{\partial^2 R_{1 1}}{\partial r_z^2}(\bar{r}_{z}) (r_{z}^*- \bar{r}_{z})^2=0, \ \ \ \mbox{Condition 2}
\label{condizio 2}
\end{eqnarray}
Since $R_{1 1} (r_z)$ is expressed in function of $f$ through Eq. (\ref{R}), this is
related to the correlation functions associated to the other directions, thus
$r_{z}^*$ is representative also for the other coordinates $x$ and $y$.
As the result, the distance from the wall of a point of TR must be always greater than $r_{z}^*$ (i.e. $r_{z}^*< y$).
Hence, it is reasonable to assume that, at the edge of the turbulent domain
\begin{eqnarray}
\begin{array}{l@{\hspace{+0.2cm}}l}
\displaystyle y_e = r_{z}^* \\\\
\displaystyle R_{T e}=15
\end{array}
\end{eqnarray}
\bigskip
In order to calculate $R_{1 1}$, $f$ is first obtained by solving
Eqs. (\ref{I2 steady})-(\ref{bc3}) which correspond
to the following Cauchy's initial condition problem (\cite{deDivitiis_3})
\begin{eqnarray}
\begin{array}{<EMAIL>}}l}
\displaystyle \frac{d f}{d \hat{r}} = F \\\\
\displaystyle \frac{d F}{d\hat{r}} = - \frac{5 f \hat{r}}{\sqrt{2(1-f)}} -
\left( \frac{1}{2} \sqrt{\frac{1-f}{2}} R_T + \frac{4}{\hat{r}} \right) F \\\\
\displaystyle f(0) = 1, \ F(0) = 0
\end{array}
\label{vk-h2}
\label{ic}
\end{eqnarray}
Hence, $R_{1 1} (r_z)$ is obtained with Eq. (\ref{R}), and $r_{z}^*$ is calculated for both the conditions (\ref{condizio 1}) and (\ref{condizio 2}).
As the result, $y^+_e$ is given by
\begin{eqnarray}
\frac{y_e}{\lambda_{T e}} = \frac{r_{z}^*}{\lambda_{T e}}
\ \ \ \mbox{where} \ R_T = R_{T e} \equiv 15
\end{eqnarray}
and $(d \lambda_T^+/d y^+)_e$ is determined according to Eqs. (\ref{ss})
\begin{eqnarray}
\displaystyle \left( \frac{d \lambda_T^+}{d y^+}\right)_e = \frac{\lambda_{T e}^+}{y_e^+}
\end{eqnarray}
\bigskip
\section{\bf Results and Discussion}
As the calculation of $k$ and $d \lambda_T^+/dy^+$ needs the knowledge
of the statistical properties of the velocity correlation,
$R_{1 1}$ is calculated for different values of $R_T$.
To this purpose, $f$ was first determined by solving numerically Eqs. (\ref{vk-h2}),
by means of the fourth-order Runge-Kutta method.
The calculation was carried out for $R_T$ = 15, 20, 40, 60 and 80, where these Reynolds numbers correspond to several distances from the wall in the turbulent region.
To obtain a good accuracy of the solutions, the step of integration is chosen to be equal to 1/40 of the estimated Kolmogorov scale $\l_K$, where
$\l_K/\lambda_T = 1/15^{1/4}/\sqrt{R_T}$ (\cite{Batchelor53}).
The results are given in Fig. \ref{figura_3} (a) which shows $f$ in terms of $r$, where the bold line represents the correlation function at the edge of TR
($R_T = 15$). This latter exhibits the ratio (integral scale)/(Taylor scale) quite similar to that of a gaussian centered in the origin, whereas in the other cases, this ratio increases with $R_T$, in agreement with the analysis of \cite{deDivitiis_3}.
A more detailed analysis about the changing of $f$ and of the corresponding energy spectrum $E(\kappa)$ with $R_T$ is reported in \cite{deDivitiis_3}.
The spanwise correlation function $R_{1 1} (r_z)$ is then calculated with
Eq. (\ref{R}), and is represented in Fig. \ref{figura_3} (b) for the same values of $R_T$.
From these data, the edge of TR, $y^+_e \equiv r_{z}^*$, is calculated for both the conditions 1 and 2.
\begin{figure}[t]
\hspace{-7.mm} \includegraphics[width=0.77\textwidth]{f_R11}
\caption{(a) Longitudinal correlation function associated to $\hat{R}_{i j}$. (b) spanwise correlation function of $u_x$, for different Taylor scale Reynolds number.
The bold lines are calculated for $R_T = 15$.}
\label{figura_3}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=0.60\textwidth]{R_T_y}
\caption{Taylor scale Reynolds number in terms of $y^+$: SL dashed lines,
TR continuous lines}
\label{figura_4}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=0.60\textwidth]{Up_y}
\caption{Dimensionless average velocity profile: SL dashed lines,
TR continuous lines}
\label{figura_5}
\end{figure}
\begin{figure}[t]
\centering
\hspace{-20.mm} \includegraphics[width=0.88\textwidth]{all}
\caption{Distribution of the dimensionless variables in the boundary layer:
SL dashed lines,
TR continuous lines.
(a) Taylor scale.
(b) Maximal Lyapunov exponent.
(c) r.m.s. of fluctuating velocity.
(d) Reynolds stress. }
\label{figura_6}
\end{figure}
The von K\'arm\'an constant and all the others parameters are shown in the table \ref{table1}.
The table reports also the values of the free parameters $A$, $B$ and $C$ associated
to the classical wall laws (\ref{wall law 00})-(\ref{wall law}), which are here calculated through the identification with Eqs. (\ref{U turbo}) and (\ref{U ss}).
In particular, $C$ is calculated with Eq. (\ref{C}), whereas $A$, which pertains the velocity law in the buffer layer, is identified as the slope $d U^+(y^+)/d \ln y^+$ of Eq. (\ref{U ss}) where $d^2 U^+(y^+)/{d \ln y^+}^2 =0$, and $B$ is consequentely determined.
The values of $k$ and $C$ are compared with the results given by different authors.
For both the conditions, $y^+_e \approx 50$ and $U^+_e \approx 15$ and this corresponds to a difference with respect to the classical data of \cite{Nikuradse} and \cite{Reichardt}, which is less than 2 $\%$.
As far as the Prandtl's length is concerned, according to Eq. (\ref{Prandtl}), it varies quite similarly to $R_T$, and its derivative, calculated for $y^+ = y_e^+$, is slightly greater than $k$
(see Eq. (\ref{Prandtl_y}))
\begin{eqnarray}
\left( \frac{d l_p^+}{dy^+} \right)_e =
1.34782... \left( \frac{d \lambda_T^+}{dy^+} \right)_e \simeq 1.00052 \ k
\end{eqnarray}
Its specific values, shown in the table, are in excellent agreement with the classical results.
For what concerns $A$, $B$ and $C$, their values agree quite well the experiments, expecially for what concerns $C$ and $A$.
Once the free parameters are identified, $R_T$
is calculated in function of $y^+$ through Eq. (\ref{R_T}) with
$O(y^+ -y^+_e)^2 \rightarrow 0$ and (\ref{ss}), for both the conditions (see Fig. \ref{figura_4}), and this corresponds to consider only the logarithmic profile of $U^+$.
Figure \ref{figura_5} shows $U^+(y^+)$ calculated with Eq. (\ref{U turbo}) and (\ref{U ss}).
It is apparent that the results agree very well with the experimental data of different authors
(\cite {Nikuradse}, \cite{Klebanoff_1}, \cite{Reichardt})
and with the formulas proposed by \cite{Spalding} and \cite{Musker}, with an error which does not never exceed 5 $\%$ in the interval of $y^+$ of the figure.
In particular, the condition 1 (Eq. (\ref{condizio 1})) provides a maximum difference
between present results and the data of Nikuradse and Reichardt, less than 3 $\%$ in
this interval.
The other variables are represented in Fig. \ref{figura_6}.
The dimensionless Taylor scale (Fig. \ref{figura_6} a), linear in the laminar sub-layer and in the buffer region, remains a rising function of $y^+$ in the turbulent zone and exhibits, there, a slope which decreases slightly with $y^+$, whereas the Lyapunov exponent (Fig. \ref{figura_6} b), defined only in TR, is represented by a monotonically decreasing function of $y^+$. This latter, being proportional to the square root of the turbulent dissipation rate ($\Lambda = u/\lambda_T$), agrees with the data of the different experiments (\cite{Fernholz}), at least where $U^+$ exhibits logarithmic profile.
\begin{figure}[t]
\centering
\hspace{-30.mm} \includegraphics[width=0.60\textwidth]{uv_zoom}
\caption{Dimensionless Reynolds stress in the laminar sub-layer
and in the buffer region}
\label{figura_7}
\end{figure}
\begin{figure}[t]
\centering
\hspace{-30.mm} \includegraphics[width=0.60\textwidth]{uvU_y}
\caption{Dimensionless rate of kinetic energy production in the laminar sub-layer
and in the buffer region}
\label{figura_8}
\end{figure}
The square root of the kinetic energy and the Reynolds stress are
shown in Fig. \ref{figura_6} (c) and (d).
These vary in TR following
Eqs. (\ref{u})-(\ref{uv}) and are monotonically rising functions of $y^+$.
As seen, $u \approx y^+$ and $\langle u v\rangle^+ \approx {y^+}^3$ in the laminar sub-layer,
whereas the Reynolds stress calculated with Eq. (\ref{uv}) in the buffer region, shows an
inflection point $y_*^+$ which represents the separation element between LL and BL.
For $y^+> y_*^+$, $-\langle u v\rangle^+$ rises with $y^+$ until to reach the turbulent region where is about constant and equal to the unity.
More in detail, Fig. \ref{figura_7} shows $-\langle u v\rangle^+$ in an enlarged region which includes LL and part of BL. The distance $y_*^+$, represented by $F_1$ and $F_2$ is $y_*^+ =$ 5.734 and 6.097, in line with the order of magnitude of the laminar sub-layer tickness, and this is achieved at about $-\langle u v \rangle^+ =$ 0.24 in the two cases.
Next, the value $-\langle u v \rangle^+$ = 0.5 is obtained in the buffer layer for $y^+ \simeq$ 10.5 and 11.2, values in very good agreement with the classical results (\cite{Tennekes}, \cite{Hinze}).
This last condition corresponds also to the maximum of the kinetic energy
production in the buffer layer, as shown in Fig. \ref{figura_8}. In this figure,
the variations of the rate of turbulent kinetic energy production due
to the mean flow are shown in LL and BL.
The shape of the diagrams follows the classical data (\cite{Tennekes}, \cite{Hinze} and references therein), and the values of $\langle u v \rangle^+$,
$\langle u v \rangle^+ dU^+/dy^+$ and of the corresponding $y^+$, are in excellent agreement with the data of \cite{Tennekes} and \cite{Hinze}.
It is worth to remark that, although the monotonic trend of $u$ and $\langle u v\rangle^+$ can contrast some experiments which can give non-monotonic variations of these variables, the values of such quantities and the corresponding $y^+$, are comparable with those of the several experiments (\cite{Fernholz}). This discrepancy can be due the fact that, here, the effects of non-homogeneity of
$\langle u_y ( p/\rho + u_j u_j/2 ) \rangle$ on the kinetic energy equation are neglected in TR, and that the average velocity is analyzed only in the logarithmic range.
\begin{table}[b]
\caption{Velocity law parameters and comparison of the results. P.R. as for "Present Result". }
\hspace{-20. mm}
\begin{tabular}{cccccccc}
Parameter & P. R. & P. R. & & & \\
& Condition 1 & Condition 2 & {Nikuradse} &
{Zagarola} & {Barenblatt} & Smith & Fernholz \\
& Eq. (\ref{condizio 1}) & Eq. (\ref{condizio 2}) & &
& & & \\
\hline
\hline \\
$\displaystyle k$ & 0.4233 & 0.3984 & 0.4 & 0.41 & 0.425 & 0.461 & 0.4 \\\\
$C$ & 6.2445 & 6.4822 & 5.5 & 5.2 & 6.79 & 7.13 & 5.1 \\\\
$A$ & 5.2384 & 5.5652 & & & & & \\\\
$B$ & -4.1226 & -4.7166 & & & & & \\\\
\hline
\hline \\
$\displaystyle \frac{r_{z}^*}{\lambda_{T e}}$ & 3.18233 & 3.38085 & & & & & \\\\
$\displaystyle \left(\frac{d \lambda_T^+}{d y^+}\right)_e$ & 0.31423 & 0.29578 & & & & \\\\
$\displaystyle \left(\frac{d l_p^+}{d y^+} \right)_e$ & 0.4235 & 0.3986 & & & & \\\\
$\displaystyle y^+_e$ & 49.3005 & 52.3760 & & & & & \\\\
$\displaystyle u^+_e$ & 0.96825 & 0.96825 & & & & & \\\\
$-\langle u v \rangle_e^+$ & 0.93750 & 0.93750 & & & & & \\\\
$U^+_e$ & 14.8250 & 15.7498 & & & & & \\\\
\hline
\end{tabular}
\label{table1}
\end{table}
\bigskip
\section{\bf Conclusions}
This work analyzes the turbulent wall laws through the Lyapunov theory of finite scale. The results, valid for fully developed flow with moderate pressure gradient,
are subjected to the hypothesis that in the turbulent region, the energy
production due to the average flow is balanced only by the dissipation rate,
whereas the non-homogeneity of $\langle u_y ( p/\rho + u_j u_j/2 ) \rangle$
is neglected.
The free parameters of the velocity law, here theoretically calculated through the
statistical properties of the velocity correlation functions, and the wall laws,
are in very good agreement with the literature.
In particular:
\begin{itemize}
\item The von K\'arm\'an constant, theoretically identified as
$k = (d R_T/d y^+)_e\approx 0.4 \div 0.42$, depends on the scale of the spanwise velocity correlation and does not requires the assumption of the logarithmic velocity profile.
\item The average velocity law and the distributions of the other dimensionless quantities such as kinetic energy and Reynolds stress in the boundary layer, agree -or at least are comparable- with experiments and direct simulations.
\end{itemize}
These results, which represent a further application of the analysis presented
in \cite{deDivitiis_3}, show that the finite scale Lyapunov analysis
can be an adequate theory to explain the wall turbulence.
\bigskip
|
\section{Introduction and outlook}
Modelling the space-time evolution of the matter created by ultra-relativistic heavy ion collisions is a great challenge. Experiments at RHIC suggest the validity of the following picture \cite{Florkowski} : (i) a large fraction of the initial kinetic energy of the colliding ions is thermalized astonishingly fast (in time $\leq$ 1 fm) forming a locally equilibrated hot and dense fireball parametrized by a profile of the hydrodynamic variables - namely the temperature, four-velocity and chemical potential fields, (ii) the strongly interacting fireball undergoes hydrodynamic expansion \footnote{This is usually modelled by the relativistic Navier-Stokes equation.}, and (iii) the initial transverse hydrodynamic flow at the time of local thermal equilibration in most cases vanishes. Most of the data at RHIC is in good agreement with this simplistic picture especially in the mid-rapidity region, i.e. for the most central collisions at the highest beam energy of $\sqrt{s_{NN}} = 200$ GeV. However, despite the success in explaining the transverse momentum spectra of hadrons, the elliptic flow coefficient, etc., this does not reproduce pion interferometric data like HBT radii leading to the well-known RHIC HBT puzzle.
It is necessary to have a better phenomenological model for the space-time evolution of the fireball to explain the data completely, and also to reduce theoretical uncertainties. The best theoretical tool at hand for studying evolution of strongly coupled matter of gauge theories in real time is the AdS/CFT correspondence. It is well-known that the AdS/CFT correspondence gives $\eta/s = 1/4\pi$ \cite{Policastro}, while the current analysis of experimental data suggests $1 < 4\pi\Big(\eta/s\Big) < 2.5$ for temperatures probed at RHIC \cite{Heinz}. In fact, the AdS/CFT correspondence also predicts systematic hydrodynamic corrections to the Navier-Stokes equation (for a review see \cite{Janik}).
Here we will propose that the AdS/CFT correspondence can be used to develop a complete phenomenology for the evolution of the strongly coupled matter, describing both the late stages of local thermalization and the subsequent hydrodynamic expansion in an unified framework. These phenomenological equations involve a closed set of equations for evolution of the energy-momentum tensor and the baryon number charge current alone.
The advantage of our proposal is that there is a very natural way to connect the expansion of the fireball with any model which describes the early stages of the collision process, as for instance the parton cascade model \cite{Geiger}. All that is needed is to match the evolution of the energy-momentum tensor and conserved charge currents before and after the matter enters in the strongly coupled phase of evolution. The entry into the strongly coupled phase can be traced through the temperature field given by the energy-momentum tensor itself as we will discuss later. Importantly, the matching with the initial regime does not require the energy-momentum tensor to be hydrodynamic.
\section{Field-theoretic grounds}
Before we use the AdS/CFT correspondence to obtain the phenomenological equations for strongly coupled irreversible processes, it will be useful to see on what field-theoretic grounds we can use only the energy-momentum tensor and conserved charge currents to construct these phenomenological equations. To keep the discussion simple, we will assume the baryon chemical potential is zero all throughout the evolution. It is in principle straightforward to include the charge currents.
It is known that the Boltzmann equation with a collision kernel determined by two body parton scattering and fragmentation processes can capture all perturbative non-equilibrium processes in non-Abelian gauge theories \cite{Arnold}. It can be shown that the relativistic semiclassical Boltzmann equation has special solutions, named "conservative solutions", and are such that they can be determined by the energy-momentum tensor alone \cite{myself1}. In this approximation, the energy-momentum tensor is parameterized by the first ten velocity moments of the parton distribution functions in phase space. It can be shown that all the higher velocity moments of the parton distribution functions have special algebraic solutions of their equations of motion, such that they are algebraic functions of the energy-momentum tensor and its derivatives, having no independent dynamics. These algebraic functions can be expanded systematically in the hydrodynamic derivative expansion and the non-hydrodynamic amplitude expansion, to be discussed later. The energy-momentum tensor follows a closed system of equations of motion, which includes the conservation of energy and momentum, but also equations for evolutions of the non-equilibrium variables giving its complete evolution. These equations can be obtained from the Boltzmann equation, and all phenomenological parameters including transport coefficients can be obtained from the collision kernel. Obviously, any solution of these equations can be lifted to a unique full solution of the Boltzmann equation.
Furthermore, any arbitrary solution of the Boltzmann equation at sufficiently late time can be approximated by an appropriate conservative solution, and any conservative solution becomes purely hydrodynamic at late time \cite{myself1}. The purely hydrodynamic solutions of Boltzmann equation are known as "normal solutions" in literature \cite{Chapman}. So, we indeed get a field-theoretic justification for using phenomenological equations involving the energy-momentum tensor alone in order to describe general irreversible processes. Also these phenomenological equations describe transition to hydrodynamic regime.
\section{General phenomenology and AdS/CFT}
In the strong coupling regime, the phenomenological equations of the evolution of the energy-momentum tensor should be obtained from gravity. Using consistent truncations of equations of motion of gravity, it can be shown that any solution of Einstein's equation with negative cosmological constant maps to a non-equilibrium state in the gauge theory via AdS/CFT correspondence, provided the solution has a regular future horizon. Furthermore, these solutions are determined uniquely by the boundary energy-momentum tensor \cite{myself2}, which by the AdS/CFT dictionary maps to the expectation value of the energy-momentum tensor in the dual state. It can be expected that when the energy-momentum tensor follows phenomenological equations with right values of the phenomenological parameters, the solutions in gravity will have regular future horizons - thus gravity should determine uniquely all phenomenological parameters.
To construct the general phenomenological equations for the energy-momentum tensor, we do not need either the Boltzmann equation or gravity, the latter are required only for determining the phenomenological parameters \cite{myself1, myself3}. Any arbitrary energy-momentum tensor can be written in the Landau-Lifshitz decomposition as :
\begin{eqnarray}\label{def}
t_{\mu\nu}(x) &=& e\left(T(x)\right) u_\mu (x) u_\nu (x) + p\left(T(x)\right) P_{\mu\nu}(x) + \pi_{\mu\nu}(x), \ \text{with}\\\nonumber \ P_{\mu\nu}(x) &=& u_\mu(x)u_\nu(x) +\eta_{\mu\nu} \ \text{and} \ u^\mu(x)\pi_{\mu\nu}(x) = 0.
\end{eqnarray}
The four-velocity $u^\mu(x)$ is the local velocity of energy-transport. The non-equilibrium part $\pi_{\mu\nu}(x)$ therefore is orthogonal to the four-velocity field and thus have six independent components; so including the velocity and temperature fields we have ten independent variables. Conformality requires $e(x) = 3p(x)$ and $\pi_{\mu\nu}(x)$ to be traceless. We will also normalize the temperature such that $e(x) = (3/4)\cdot \left(\pi T(x)\right)^4$.
The constraints of Einstein's equations automatically gives the conservation of energy and momentum :
\begin{equation}\label{cons}
\partial^\mu t_{\mu\nu} = 0.
\end{equation}
Without loss of generality, $\pi_{\mu\nu}$ can be split into a purely hydrodynamic part $\pi_{\mu\nu}^{(h)}$ and a non-hydrodynamic part $\pi_{\mu\nu}^{(nh)}$ which cannot be determined by hydrodynamic variables alone. Thus,
\begin{equation}\label{spl}
\pi_{\mu\nu} = \pi_{\mu\nu}^{(h)}+ \pi_{\mu\nu}^{(nh)}.
\end{equation}
The hydrodynamic part, $\pi_{\mu\nu}^{(h)}$ has a purely hydrodynamic derivative expansion, the expansion parameter $\epsilon$ is the ratio of the typical length scale of variation to the mean-free path. Requiring conformal invariance, and using the AdS/CFT correspondence to obtain the transport coefficients, we get up to second order in derivative expansion \cite{Janik},
\begin{eqnarray}\label{soh}
\pi_{\mu \nu}^{(h)} &=& - 2(\pi T)^3 \sigma_{\mu \nu} +
(2 - \ln 2)(\pi T)^2 \mathcal{D}\sigma_{\mu \nu} + 2 (\pi T)^2
\left(\sigma_{\mu}^{\phantom{\mu}\alpha}\sigma_{\alpha\nu} -
\frac{1}{3}P_{\mu \nu} \sigma_{\alpha \beta}\sigma^{\alpha
\beta}\right) \\\nonumber &&+\ln 2 (\pi T)^2
(\sigma_{\mu}^{\phantom{\mu}\alpha}\omega_{\alpha\nu}
+\sigma_{\nu}^{\phantom{\nu}\alpha}\omega_{\alpha\mu} ) +
O(\epsilon^3),
\end{eqnarray}
where $\sigma_{\mu\nu}$ is the shear-stress tensor, $\omega_{\mu\nu}$ is the velocity-vortex, and $\mathcal{D}$ is the Weyl-covariant convective derivative.
The non-hydrodynamic part $\pi_{\mu\nu}^{(nh)}$ has an additional amplitude parameter $\delta$, which is the ratio of the typical non-hydrodynamic shear-stress to the equilibrium pressure. However, unlike the hydrodynamic variables, in a local inertial frame where the energy flow vanishes, close to equilibrium $\pi_{\mu\nu}^{(nh)}$ is slowly varyiing in space but not in time. So, at every order in amplitude expansion we must sum over all time-derivatives, or to state in a Lorentz and Weyl covariant manner - all Weyl-covariant convective derivatives $\mathcal{D}$. Furthermore, it should be possible to set consistently $\pi_{\mu\nu}^{(nh)}$ to zero as we know that the purely hydrodynamic sector exists in both the Boltzmann equation and gravity. Putting all these requirements together, and expanding both in $\epsilon$ and $\delta$
we get the most general Weyl and Lorentz covariant phenomenological equation for $\pi_{\mu\nu}^{(nh)}$ \cite{myself3}:
\begin{eqnarray}\label{rcg}
\left(\displaystyle\sum\limits_{n=0}^{\infty}D_{R}^{(1,n)}(\pi T)^n
\mathcal{D}^n\right)\pi_{\mu \nu}^{(nh)} &=&
\frac{(\pi T)\lambda_1 }{2}
\left(\pi_{\mu}^{(nh)\alpha}\sigma_{\alpha\nu}
+\pi_{\nu}^{(nh)\alpha}\sigma_{\alpha\mu}- \frac{2}{3}P_{\mu
\nu} \pi_{\alpha \beta}^{(nh)}\sigma^{\alpha \beta}\right)
\\\nonumber
&&+\frac{(\pi T)\lambda_2 }{2}
\left(\pi_{\mu}^{(nh)\alpha}\omega_{\alpha\nu}
+\pi_{\nu}^{(nh)\alpha}\omega_{\alpha\mu}\right) \\\nonumber
&&- (\pi T)^{4}\displaystyle\sum\limits_{n=0}^{\infty}\displaystyle\sum\limits_{\substack{m=0 \\ n+m \ is \ even}}^{n}D_R^{(2,n,m)}(\pi T)^{n}\displaystyle\sum\limits_{\substack{a,b=0 \\ a+b=n \\ |a-b|=m}}^{n}\Bigg[\mathcal{D}^a
\pi_{\mu}^{(nh)\alpha}\mathcal{D}^b
\pi_{\alpha\nu}^{(nh)}\\\nonumber &&\qquad\qquad\qquad\qquad\qquad\qquad\qquad -\frac{1}{3}P_{\mu\nu} \ \mathcal{D}^{a}
\pi_{\alpha\beta}^{(nh)}\mathcal{D}^b \pi^{(nh)\alpha\beta}\Bigg]\\\nonumber
&&+ O(\epsilon^2\delta, \epsilon\delta^2 ,\delta^3).
\end{eqnarray}
It is very hard to give a general proof of validity of this equation in gravity, nevertheless it has been shown that it reproduces the gravity solutions dual to homogeneous relaxation \cite{myself3}. In such cases, the hydrodynamic variables are constant in space and time, while $\pi_{\mu\nu}^{(nh)}$ is spatially homogeneous. Such solutions also exist in the Boltzmann equation. It describes well the transition to local equilibrium. It has been found that the future horizon is regular, provided \emph{all convective derivatives are summed over at each order in the amplitude expansion} as expected. Furthermore, we obtain a complicated recursion relation for the phenomenological parameters $D_{R}^{(1,n)}$ and $D_R^{(2,n,m)}$, the first few terms being \cite{myself3}:
\begin{equation}
D_R^{(1,0)}= -1, \ \ D_R^{(1,1)}=-(\pi/2) - (1/4) \ \ln 2, \ \ \text{etc.;} \quad D_R^{(2,0,0)} = 1/2, \text{etc.}
\end{equation}
In order to obtain other parameters in (\ref{rcg}) like $\lambda_1$, $\lambda_2$, etc. it will be necessary to consider more general inhomogeneous configurations. Proving that the eqs. (\ref{def}), (\ref{cons}), (\ref{spl}), (\ref{soh}) and (\ref{rcg}) give all solutions of pure gravity in AdS with regular future horizons for right phenomenological parameters, which have been partially determined here, will give us further confidence in these phenomenological equations \footnote{In practice, one needs to use these equations in a coordinate system better adapted for the late equilibrium state of the fireball which is an ideal fluid undergoing boost-invariant expansion. This coordinate system comprises of the proper time coordinate $\tau$ of the late time expansion, the coordinate $y$ parameterizing rapidity, and the two transverse coordinates $x^1$ and $x^2$. } .
|
\section{Tools from Non-equilibrium Field Theory in Molecular QED}
The interaction between electrically neutral yet polarizable particles,
commonly named after J. D. van der Waals, F. London, H. Casimir, and D. Polder,
is one of the fundamental problems in atomic and molecular physics.
While the two-body potential between two particles (atoms, molecules, nanoparticles)
in their internal ground state
is unambiguous\cite{Casimir_1948,Power_1983,McLachlan_1963,Craig_1998}, apparently incompatible results have been obtained
if one of the atoms is prepared in an excited state.
The quantity of interest is the long-range part of the potential
that overwhelms the familiar van der Waals interaction:
some calculations found this part to oscillate spatially%
\cite{McLone_1965,Philpott_1966,Kweon_1993},
while later work found a monotonic power law%
\cite{Power_1993}\nocite{Power_1995,Sherkunov_2007,Sherkunov_2009}.
It is the aim of the present paper to understand these differences\cite{Power_1995,Sherkunov_2007} better.
We handle this non-equilibrium problem with a diagrammatic (Feynman graph) expansion.
The starting point of our quantum mechanical description is the electric dipole coupling Hamiltonian
\begin{align}
H_{AF}
=
-
\sum_{n} E_i(x_n)
\left[d_i^{n}\psi^e(x_n) \psi^{g\dagger}(x_n)
+ d_i^{n *}\psi^g(x_n) \psi^{e\dagger}(x_n)\right]
~.
\label{eq:dipole-coupling-Hamiltonian}
\end{align}
For simplicity, we treat the atoms as pointlike two-level objects located
at the spacetime coordinates $x_n = \{\V{r}_n, t_n\}$.
The fermionic operators $\psi^{a\dagger}(x)$ create atoms in state
$a= g, e$ at $x$. Assuming that the transition dipoles $d_i^n$ are
real, we work in the following with an effectively scalar electric field
$E( x ) = d_i^n E_i ( x )$ whose space argument accounts for possible
differences in the transition dipoles.
The textbook treatment of dispersive atom-atom
interactions\cite{Craig_1998,Salam_2009}
employs stationary fourth-order perturbation theory in $H_{AF}$.
The summation over all possible intermediate states leads to a rather large
number of terms.
In this work, we discuss a more concise treatment
using the closed time-path contour formalism due to Schwinger, Craig, Mills,
and Keldysh (see Ref.\,\refcite{Danielewicz_1984}).
Following Sherkunov
\cite{Sherkunov_2007,Sherkunov_2009,Sherkunov_2005},
the level shift of a ground-state atom is extracted from its full propagator
\chdid{[$T$ order symbol removed]} \todo{HH: \Large$\checkmark$}
\begin{align}
i g^{\rm (full)}_{\alpha \beta}(x,y)
&\equiv
\langle
T_c \bigl\{
S_2(-\infty, \infty)
S _1(\infty, -\infty)
\psi^g_\alpha (x) \psi^{g \dagger}_\beta (y)
\bigr\}
\rangle^{\rm conn}
~.
\label{eqn:Dyson_Feynman}
\end{align}
The greek subscripts take the values 1 or 2,
denoting the branch of the Keldysh time contour
which extends from $t=-\infty$ to $\infty$ (branch 1) and back (branch 2).
The prescription $T_c\{ \dots
\}$ orders operators according to their time arguments on the contour.
Finally,
$S$ is the standard scattering operator,
the brackets $\langle \dots \rangle$ denote
an eigenstate of the non-interacting theory%
\chdid{[CH: need a more general state here, since the excited atom A will
come in]} \todo{HH: \Large $\checkmark$}%
,
and only \emph{connected} products contribute.
The Feynman formalism\cite{Schiefele_2010,Schiefele_2011} is valid for known in- and out-states,
which is the typical scenario
in equilibrium systems at zero temperature.
If, however, a general initial state
(including, in our case, unstable excited atomic states)
is prepared and then left to evolve under the interaction,
the Keldysh formalism is necessary to describe all energetically allowed processes.
In the following, a propagator with both times on the forward
branch will be used. Its series expansion is from
Eq.\,(\ref{eqn:Dyson_Feynman})
\chdid{[prefactor $-1$ added! not $c$-integration!]}\todo{HH:\Large $\checkmark$}
\begin{align}
& i g^{\rm (full)}_{1 1}(x,y)
=
\langle
T \bigl\{
\psi^g (x) \psi^{g \dagger} (y)
\bigr\}
\rangle
\nonumber
\\
&-
\frac{1}{2}
\int d{t_1} d{t_2} \,
\langle
T_c \bigl\{
\sum_{\alpha,\beta = 1}^2\,
(-1)^{\alpha + \beta}
H_{AF, \alpha} (t_1) H_{AF, \beta}(t_2)
\psi^g_1 (x) \psi^{g \dagger}_1 (y)
\bigr\}
\rangle^{\rm conn}
+
\dots
\;.
\label{eq:Dyson-Keldysh-11}
\end{align}
where the first line is the ordinary (bare) Feynman propagator given
in Eq.\,(\ref{eq:g-propagator}), with $T$ being the standard time ordering.
We will see below that processes involving only ground-state atoms can be entirely
described in terms of Feynman propagators.
The next section presents the calculation of self-energies
for atoms and photons, and recovers the energy shifts and decay widths in a
system of two particles. A comparison of the Feynman and Keldysh results
suggests an interpretation of the apparent disagreement of earlier
results: these may apply in different time regimes after preparation, separated
by the time scale that characterizes energy transfer by an exchange of
resonant photons.
Finally, we discuss how resonant long-range potentials might influence some systems of biological relevance that are not commonly considered in the community present at this conference.
\section{Self-Energies with Dressed Photons}
Our perturbative analysis is built from two basic expressions:
the self-energies of an atom coupled to the photon field
and of a photon coupled to a (second) atom, respectively.
Conceptually, this is closely related to the notion of dressed states\cite{Compagno_1995}.
The real and imaginary parts of the self-energy
\chdid{[CH: minus sign in decay rate]}
\begin{align}
\Sigma_{1 1} = \Delta E - i \frac{ \Gamma }{2}
\label{eqn:sigma_re_im}
\end{align}
can be identified as the energy-shift and the inverse lifetime (decay width) of the particle.
We first calculate the atomic self-energy at next-to-leading order in the
coupling to a generic photon field.
In a second step, the interatomic interaction is identified by ``dressing'' the photons with a second atom.
We only have to consider terms where photons are connecting the two atoms,
provided we assume the transition frequencies to include the single-particle
Lamb-shift.
\subsection{Single atom plus field: atom self-energy}
\label{sec:atom-is-dressed}
We start by considering the propagator of a ground-state atom $B$
(transition frequency $\omega^B$)
at the one-loop level.
Using the Feynman rules from the Appendices and App. C of
Ref.\,\refcite{Schiefele_2011}, the evaluation of
Eq.\,(\ref{eq:Dyson-Keldysh-11}) yields
\chdid{[CH: factor $i$ removed, from expansion of (\ref{eq:g-and-e-propagator-11})
around $\epsilon_g$;
consistent with Eq.\,(22) of Sherkunov2007]}
\begin{align}
g_{1 1}^{(2)} &(x,x')
=
\GPropOneLoop{-0.ex}{0.25\textwidth}
\nonumber
\\
&=
\int d^3 x_1
\int d^3 x_2
\int d \omega
\, e^{-i\omega(t - t')}
g_{1 1} (\omega, \V{r},\V{r}_1)
\Sigma_{11}^{g} (\omega, \V{r}_1,\V{r}_2)
g_{1 1} (\omega, \V{r}_2,\V{r}')
\;,
\label{eqn:def_selfenergy}
\end{align}
where the self energy for a pointlike atom is to lowest order
\begin{align}
\Sigma_{1 1}^{g B} ( \omega_{\rm in} )
&=
\SigmaG{-2ex}{0.17\columnwidth}
=
-i
\int
\frac{d\omega}{2\pi} \,
D_{1 1}(\omega, \V r_B, \V r_B)
e_{1 1}( \omega_{\rm in} + \omega )
\;.
\label{eqn:Sigmag}
\end{align}
The excited state propagator $e_{1 1}$ is given in
Eq.\,(\ref{eq:g-and-e-propagator-11}) and $D_{1 1}$ denotes the Feynman
propagator for photons. For the self-energy
$\Sigma^e_{1 1}$ of an excited atom, replace $e_{1 1}$ by
$g_{1 1}$ in the integral.
This result can be brought into a more familiar form (see
Ref.\,\refcite{Wylie_1985}) by evaluating it
on the mass shell, $\omega_{\rm in} = \epsilon^B_g$, and writing the
integral over positive frequencies only (recall that
$D_{1 1}(\omega)$ is even in $\omega$):
\begin{align}
\Sigma^{g B}_{1 1}( \epsilon^B_g )
&=
i
\int_0^\infty
\frac{d\omega}{2\pi} \,
D_{1 1}(\omega, \V r_B, \V r_B)
\alpha_{1 1}^{g B} (\omega)
\label{eqn:sigmaGp}
\;.
\end{align}
Here we recover the Feynman-ordered polarizability
\begin{align}
\alpha_{1 1}^{g B} (\omega)
&=
\frac 1 {\omega^{B} - \omega - i\, 0^+}
+
\frac 1 {\omega^{B} + \omega - i\, 0^+}
~,
\label{eqn:alpha11}
\end{align}
that is evaluated explicitly in \ref{sec:polarizations}.
(For the excited two-level atom $A^*$, replace $\omega^B \to -\omega^A$.)
Note that the propagators here
have poles in the upper left and lower right quadrant of the complex
frequency plane
(Feynman prescription). They coincide, however, with the usual retarded
response functions at positive frequencies, and only these appear in
the complex self-energy~(\ref{eqn:sigmaGp}).
Substituting the bare photon propagator
into $\Sigma^{g/e}_{1 1}$
yields, according to \pr{eqn:sigma_re_im}, an infinite energy shift which is,
of course, the unrenormalized Lamb shift of atom $B$.
The imaginary part of $\Sigma^{g}_{1 1}$ is nonzero if the temperature
$T > 0$ [see Eq.\,(\ref{eq:photon-11-finite-T})], describing the absorption rate
of thermal photons. At $T = 0$, an imaginary part is found for the excited
atom $A^*$: this comes from
a pole of $\alpha^{e A}_{1 1}( \omega )$ in the upper right quadrant and
gives the spontaneous decay rate
$\Gamma^{e A}_0 = 2 \,{\rm Im}\, D_{1 1}(\omega_A, \V r_A, \V r_A ).$
\subsection{Photon propagation in the presence of a second atom}
\label{sec:NLO_photon}
To evaluate the impact of a second atom (labeled $A$) on the self-energy of
atom $B$, we take the diagram~(\ref{eqn:Sigmag}) and replace the
photon propagator by its next-to-leading order correction:
$$
\SimplePhoton{-0.1ex}{0.08\textwidth}
\;\to\;
\DressedPhotonG{-1.ex}{0.15\textwidth}
{\rm or}
\DressedPhotonE{-1.ex}{0.15\textwidth}
~,
$$
depending on the state of atom $A$. This ``modular'' strategy is, of
course, consistent with the fourth-order expansion of the
atom propagator $g_{1 1}$ in Eq.\,(\ref{eqn:Dyson_Feynman}).
\todo{[CH: should be true, but did somebody check? HH: That is how we compared to the ear-diagrams, so: \Large $\checkmark$]}
The bare photon propagator $D_{1 1}( x, x' )$ is given in
\ref{sec:free_photons}, and its correction to leading order
recovers Eq.~(28) of Ref.\,\refcite{Sherkunov_2007}
\todo{HH: Do we really need this reference?}
\begin{align}
D^{(2)a}_{1 1} (x', x)
&=
\frac{- i}{2}
\int\! d t_1 d t_2 \,
\langle T_c \bigl\{
\sum_{l, l' = 1}^2
(-1)^{l+l'}
H_{AF,l}(t_1) H_{AF,l'}(t_2)
E_1(x') E_1(x)
\bigr\}
\rangle_{ a }^{\rm conn}~.
\nonumber
\end{align}
where $a = e, g$ denotes the state of atom $A$.
For the purpose of our analysis it is sufficient to restrict the field to zero temperature
which simplifies the treatment in the frequency domain considerably
(see~\ref{sec:free_photons}).
We find from Wick's theorem\cite{Danielewicz_1984}
\todo{[CH: please check space arguments and minus sign. HH: Looks good \Large$\checkmark$]}
\begin{eqnarray}
D^{ (2) a }_{1 1}(\omega > 0, \V r', \V r)
&= & D_{1 1}(\omega, \V r', \V r_A ) \alpha^{ a A }_{1 1}(\omega)
D_{1 1}(\omega, \V r_A, \V r )
\nonumber\\
&& {}
- D_{1 1}(\omega, \V r', \V r_A) \alpha^{ a A }_{1 2}(\omega)
D_{2 1}(\omega, \V r_A, \V r)
\label{eq:D11-dressed-photon}
\end{eqnarray}
where the atomic polarizabilities are given in \pr{sec:polarizations}.
Note that this contains an extra term that is not in the form of Feynman
propagators.
There is no connected diagram involving $D_{2 2}$, and
at positive frequencies, $D_{1 2}(\omega) = 0$
(see~\ref{sec:free_photons}).
\subsection{Two ground-state atoms}
If atom $A$ is in the ground state ($a = g$), the contribution from the
second Keldysh branch vanishes because $\alpha^{g A}_{1 2}( \omega > 0 )
= 0$.
We obtain a two-atom self-energy by substituting the dressed photon
propagator $D^{(2)g}_{1 1}$ into $\Sigma^{g}_{1 1}$,
\todo{[CH: factor $i$ added, please check. HH: Is this a global prefactor for all self-energies?]}
\begin{align}
\Sigma^{gg}_{1 1}
&=
\BoxGG{-2.5ex}{0.25\columnwidth}
+
\CrossGG{-2.5ex}{0.25\columnwidth}
\nonumber
\\
&=
i
\int_0^\infty \frac{d \omega}{2 \pi }\alpha^{g B}_{11}(\omega)
\alpha^{g A}_{1 1}(\omega)
D_{1 1}(\omega, \V r_A, \V r_B)
D_{1 1}(\omega, \V r_B, \V r_A)
~.
\nonumber
\end{align}
Only Feynman-ordered quantities contribute, and indeed, this situation corresponds
to the ground state of the non-interacting theory where no non-equilibrium
formalism is needed.
The integration lends itself to a rotation to imaginary frequencies, and one
sees that $\Sigma^{g g}_{1 1} = \Delta E^{g g }$ is a purely real
two-body potential, equal to the well-known Casimir-Polder potential if the
two atoms are embedded in free space.
Note that the same calculation in traditional perturbation theory, e.g. in
Ref.\,\refcite{Craig_1998},
involves 12 diagrams with time-directed photon-lines rather than the two
Feynman diagrams above.
\subsection{Ground-state atom and excited atom}
The additional term in the photon propagator~(\ref{eq:D11-dressed-photon})
makes the Feynman and Keldysh results differ. We find the following
self-energy for the ground-state atom $B$ in the presence of an
excited atom $A^*$
\js{Probably we don't need to put labels on the points here.}
\todo{[CH: factor $i$ added and position arguments corrected, please check.]}
\begin{align}
\Sigma^{ge}_{1 1}
&=& i \int_0^\infty \frac{d \omega}{2 \pi }
\alpha^{g B}_{11}(\omega)
\left\{\right
D_{1 1}(\omega, \V r_B, \V r_A)
\alpha^{e A}_{1 1}( \omega )
D_{1 1}(\omega, \V r_A, \V r_B) &&
\nonumber
\\
&& \left. {}- D_{1 1}(\omega, \V r_B, \V r_A)
\alpha^{e A}_{1 2}( \omega )
D_{2 1}(\omega, \V r_A, \V r_B)
\right\}&&
\label{eqn:dressed_photon_2}
\\
\nonumber%
&=& \BoxStandard{-2.5ex}{0.25\columnwidth} + \CrossStandard{-2.5ex}{0.25\columnwidth}\nonumber&&\\
&& +\BoxRightEar{-2.5ex}{0.25\columnwidth} + \CrossRightEar{-2.5ex}{0.25\columnwidth}&&
\label{eqn:dressed_photon_2a}
\end{align}
The first line is the result obtained in Refs.\,\refcite{McLone_1965}--\refcite{Kweon_1993}.
In our formalism, this part of the result contains only time-ordered (Feynman) quantities and is represented
by the first two diagrams of \pr{eqn:dressed_photon_2a}.
We will argue in the next section under which conditions this line gives already the full result.
The second set of diagrams illustrates the terms that arise from
the second Keldysh branch. There is one vertex beyond the (thick red) bars
that corresponds to an interaction operator $H_{AF,2}( x )$
on the second Keldysh branch. It will end leftmost after the contour ordering
$T_c$ and can be interpreted as acting directly on the outgoing states.
This modification of the out-state translates the equilibration of the
initial state (prepared at $t = - \infty$ and left to evolve under the interaction).
The thick red bar thus illustrates
the split between the branches of the Keldysh contour: it corresponds, apart
from a prefactor,
to cutting the diagrams and putting the loose ends on the mass-shell.
The resulting two off-diagonal processes conserve energy (tree structure
of the diagrams)
and can be read as
spontaneous emission from atom $B$ after a resonant photon exchange (left
diagram) and as emission from atom B stimulated by a photon from atom A
(right diagram).
To provide a more transparent physical interpretation,
we bring the above result into a form involving retarded and advanced
response functions.
These are related to the Keldysh propagators by
$\alpha_{11} = \alpha_R + \alpha_{1 2}$ \todo{HH: Changed sign \Large$\checkmark$}
and $D_{2 1} = D_{1 1} - D_A$
(see Refs.\refcite{Danielewicz_1984,vanLeeuwen_2006}). Remembering that
for positive frequencies and $T = 0$, we have
$D_{1 1}( \omega ) = D_R( \omega )$ and
$\alpha^g_{1 1}( \omega ) = \alpha^g_{R}( \omega )$,
we get
\chdid{[CH: factor $i$ added, plus sign corrected in last line]}
\todo{HH: Swapped back signs of middle lines}
\begin{align}
\Sigma^{ge}_{1 1}
= i \int_0^\infty \frac{d \omega}{2 \pi }
\alpha^{g B}_{R}(\omega)
\left\{ \right
& D_{R}(\omega, \V r_B, \V r_A) \alpha^{e A}_{R}( \omega ) D_{R}(\omega, \V r_A, \V r_B)
\nonumber\\
& {} + D_{R}(\omega, \V r_B, \V r_A) \alpha^{e A}_{1 2}( \omega ) D_{R}(\omega, \V r_A, \V r_B)
&(*)
\nonumber\\
&{} - D_{R}(\omega, \V r_B, \V r_A) \alpha^{e A}_{1 2}( \omega ) D_{R}(\omega, \V r_A, \V r_B)
&(*)
\nonumber\\
& {} + D_{R}(\omega, \V r_B, \V r_A) \alpha^{e A}_{1 2}( \omega )
D_{A}(\omega, \V r_A, \V r_B)
\label{eqn:dressed_photon_3}
\left. \right\}
~.
\end{align}
The two lines indicated by the asterisk $(*)$ cancel out exactly.
The first line gives an integrand regular in the upper right quadrant and
can be rotated onto the imaginary axis; it does not yield any oscillating
contributions. Within the two-level approximation, one has
$\alpha_{R}^e = -\alpha_{R}^g$ [see Eq.\,(\ref{eqn:alpha11})], so that
the first term is equal, up to a global sign, to the ground-ground
self-energy $\Sigma^{gg}_{1 1}$.
The last line gives a purely resonant contribution because
$\alpha^{e A}_{1 2}( \omega ) = 2\pi i \, \delta( \omega - \omega_A )$
according to \pr{eqn:pi_12}. It involves the modulus squared of the
(retarded) photon Green function because $D_A( \omega ) = D_R^*( \omega )$
at real frequencies. We thus rewrite Eqn.\eqref{eqn:dressed_photon_3} as
\todo{[CH: minus sign added in last term, consistent with imaginary
part; check real part.]}
\begin{align}
&\Sigma^{ge}_{1 1}
= - \Sigma^{gg}_{1 1}
-
\alpha^{g B}_{R}(\omega_A) \,
|D_{R}(\omega_A, \V r_B, \V r_A)|^2
\label{eq:eg-potential-Keldysh}
\\
&=
{} - \hspace{-.2cm}\BoxGG{-2.5 ex}{.25\columnwidth} \hspace{-.25cm}
- \hspace{-.2cm}\CrossGG{-2.5ex}{.25\columnwidth}
\hspace{-.25cm}
- \left|\ResonanceInteraction{-2.5ex}{.2\columnwidth}\right|^2
\alpha^{g B}_{1 1}(\omega_A)
~,
\nonumber
\end{align}
where the last term can be identified as F\"orster resonant energy
transfer\cite{Power_1983,Salam_2009,Forster_1949,Andrews_1989,Cohen_2003,Cohen_2003a} (FRET).
Rather than distinguishing absorption and emission,
the above reordering has led to a separation of dispersion (nonresonant
or virtual photons) and FRET effects.
The resonant energy exchange makes the ground-state atom $B$ unstable,
and we recover the rate for FRET
\begin{equation}
\Gamma_{\rm FRET} = - 2
\, {\rm Im} \bigl[ \Sigma^{g e}_{1 1} \bigr]
=
2 \pi \delta(\omega_B - \omega_A)
|D_{R}(\omega_A, \V r_B, \V r_A)|^2
~,
\nonumber
\end{equation}
in full agreement with the Golden Rule. The
generalization of this rate to molecular emission and absorption spectra
of finite width is straightforward and well-known \todo{good book to quote?}.
The FRET process comes along with a resonant two-body potential (the
real part of $\Sigma^{ge}_{1 1}$) that has been discussed in
Refs. \refcite{Power_1993}--\refcite{Sherkunov_2009}. This potential
does not oscillate spatially and lends itself to a simple semi-classical
picture: the polarization energy of atom $B$ in the time-averaged field
of a dipole source located at atom $A$, whose amplitude is fixed by the
transition dipole\cite{Power_1995}.
\chdid{this picture confirms the sign in Eq.\,(\ref{eq:eg-potential-Keldysh})
because the polarization energy is $- \alpha |E^2|$}
\subsection{Transient spatial oscillations of the resonant potential}
%
It is instructive to evaluate the contribution that was cancelled from
Eq.\,(\ref{eqn:dressed_photon_3}) (upper line marked (*)).
The explicit form of $\alpha^{eA}_{1 2}$ from \pr{eqn:pi_12}
results in a spatially oscillating contribution to the self-energy
\chdid{[CH: only one sign]}\todo{HH: chose $-$}
\begin{align}
\Delta \Sigma^{eg}_{1 1} = - \alpha^{g B}(\omega_A)
D_{1 1}(\omega_A, \V r_B, \V r_A)
D_{1 1}(\omega_A, \V r_A, \V r_B)
~.
\label{eq:extra-term-with-oscillations}
\end{align}
Hence the first two (Feynman) diagrams in Eq.\,(\ref{eqn:dressed_photon_2})
give
\chdid{[CH: sign changed]}
\begin{align}
\BoxStandard{-2.5ex}{0.25\columnwidth} + \CrossStandard{-2.5ex}{0.25\columnwidth}
= - \Sigma_{1 1}^{gg} + \Delta \Sigma_{1 1}^{eg}
~.
\label{eqn:Feynman_result}
\end{align}
Since the potential $\Sigma_{1 1}^{gg}$ is monotonous, the extra
term~(\ref{eq:extra-term-with-oscillations}) is responsible for
the spatially oscillating potential
found in Refs.\,\refcite{McLone_1965}--\refcite{Kweon_1993}
using equilibrium (Feynman) theory.
If all four diagrams are taken into account, one reaches a fully equilibrated
state and the interaction $\Sigma_{1 1}^{eg}$
{}[Eq.\,(\ref{eq:eg-potential-Keldysh})]
does no longer contain any spatial oscillations.
This is the result in Power and Thirunamachandran's
calculations\cite{Power_1993,Power_1995}.
\todo{Did Sherkunov's 2005 paper give the oscillating result?}
\todo{What did the erratum of Kweon \& Lawandy from 1994 find? They
are more general because they take states above $|e\rangle$.} \todo{HH: To me it looks as if both groups (incl. the Erratum) found oscillations, can someone confirm?}
We suggest the following scenario to understand the physical relevance of
the two results, using the narrative of atom dressing\cite{Compagno_1995}.
The interchange of virtual (or non-resonant) photons responsible for
the monotonous part of the potential is very fast. This short-time behaviour
is similar for both ground-state and excited atoms. The exchange of
near-resonant photons takes a longer time due to the small frequency
differences involved, and it leads to the equilibration of the initial state.
This happens on the scale $1/\Gamma_{\rm FRET}$ set by the F\"orster rate.
This time appears as a lower limit for the validity of the equilibrated potential
obtained from $\Sigma_{1 1}^{e g}$. An upper limit is set by the spontaneous
lifetime $1/\Gamma^e_0$ of the atom $A^*$
when photons appear in free space modes\cite{Sherkunov_2009,Cohen_2003}. The two time scales are well
separated at short distance (non-retarded range, typical for FRET
experiments).
The spatial oscillations of the potential~(\ref{eqn:Feynman_result})
may therefore appear as a transient effect right after the preparation.
We speculate that they remain visible if the excited state is
continuously replenished with a sufficiently high rate (larger than
$\Gamma_{\rm FRET}$).
This would require a time-dependent perturbative calculation similar to
Ref.\refcite{Power_1983} that is beyond
the scope of this paper.
We note that a spatially oscillating dispersion interaction has been found at
short times
in the somewhat analogous situation of an excited molecule close to a
surface\cite{Ellingsen_2009}.
\section{Discussion and Conclusions}
The apparently differing predictions for the resonant part of the two-body potential between a ground-state and an excited atom have been a source of confusion for quite some time. We have reviewed this system in a non-equilibrium description (Keldysh closed time-path formalism) and identified how contributions that oscillate spatially disappear in such a description.
The two results may correspond to different physical setups or to situations separated by a characteristic time scale, which we related to the rate of energy transfer (FRET): the oscillatory potential
(Refs.\,\refcite{McLone_1965}--\refcite{Kweon_1993}) holding on very short time scales, and eventually evolving into the non-oscillatory one of Refs.\,\refcite{Power_1993}--\refcite{Sherkunov_2009}.
In molecular physics, F\"orster broadening of atomic or molecular spectra is well known, but there seems to be less literature on the forces that come along with such resonant exchanges of energy\cite{Malnev_1970}.
Actually, the term \emph{F\"orster force} was coined as late as 2003 by Cohen and Mukamel\cite{Cohen_2003,Cohen_2003a}.
These forces may, however, play an important role in systems that stand in the focus of recent research,
ranging from cold atoms\cite{Anderson_1998,Mourachko_1998} to quantum dots, NV centers, and even biomolecules such as proteins and DNA.
In the latter context, it has been proposed \todo{HH: I don't know by whom!}
that diffusion-limited reactions between an excited and a ground-state reactant are facilitated by a state-selective force. The force would modify the motion of the molecules relative to undirected Brownian diffusion which may help to understand unusually high reaction rates\cite{Berg_1985}.
In biological systems in particular, the excited state may be pumped by an energy source (photoabsorption, chemical). Then the energy flow through the system will determine to what extent the scenario involving equilibration after excitation actually applies.
Such questions certainly deserve further investigation, and underline that talking about unstable states as in dispersion interactions between atoms or molecules requires a careful characterization of the \emph{excited state} in question.
We thank R. Behunin, S. Buhmann, G. Cammarata, F. Intravaia, V. Mkrtchian,
R. Passante, G. Pieplow, J. Preto, S. Scheel, and S. Spagnolo for helpful discussions.
Partial financial support by DFG, ERC, GIF, and DAAD is acknowledged.
Participation in QFExt11 was made possible by ERC and DAAD.
All diagrams were created using Jaxodraw\cite{Binosi_2004}.
\begin{appendix}
\section{Atom fields and propagators}
\label{sec:polarizations}
Atoms are considered distinguishable by an index $n = A, B$, with an internal two-level structure of energies $\epsilon_{g,e}^{n}$.
The results will only depend on the Bohr frequencies
$\omega^n \equiv \epsilon_e^n - \epsilon_g^n$~.
The annihilation operator $\psi_a^n( x )$ evolves in the interaction picture
proportional to $\exp( -i \epsilon_{a}^{n} t )$ ($a = g, e$).
Atoms are considered immobile and pointlike so that the position
dependence is carried by the field operator
in the interaction
Hamiltonian~(\ref{eq:dipole-coupling-Hamiltonian}).
The bare (leading order) propagator for a ground-state atom is given by
\begin{align}
g_{1 1}^{n} ( x', x )
&=
-i \langle \psi_g^{n} (x') \psi_g^{n \dagger} (x) \rangle
\theta(t' - t)
&=
\int \frac{d \omega}{2 \pi} \,
\frac{e^{-i \omega (t' - t)}}
{\omega - \epsilon_g^{n} + i\, 0^+}
\;.
\label{eq:g-propagator}
\end{align}
For an excited atom, $e_{1 1}^n$ is obtained using $\epsilon_e^n$ in place
of $\epsilon_g^n$.
The step function $\theta( t' - t )$ arises because in this non-relativistic
theory, there are no anti-particles.
In the frequency representation, we thus have the Feynman rules
\todo{harmonize diagrams}
\begin{align}
g_{1 1}(\omega) = \frac{1}{\omega - \epsilon_g + i\, 0^+} =
\includegraphics*[width=0.1\textwidth]{diagrams/ground}%
\;, \quad
e_{1 1}(\omega) = \frac{1}{\omega - \epsilon_e + i\, 0^+} =
\raisebox{0ex}{\includegraphics*[width=0.1\textwidth]{diagrams/eprop1}}
\label{eq:g-and-e-propagator-11}
\end{align}
We also need Keldysh propagators for the atomic polarization operator
$P( x_n ) \equiv \psi^e(x_n) \psi^{g\dagger}(x_n)
+ \psi^g(x_n) \psi^{e\dagger}(x_n)$ that appears in the dipole
interaction~(\ref{eq:dipole-coupling-Hamiltonian}).
The general correlation is
\begin{equation}
\alpha^{an}_{\alpha\beta}( x', x ) =
i \langle T_c\{ P_\alpha( x'_n ) P_\beta( x_n ) \}\rangle_a
\label{eq:def-polarization-correlation}
\end{equation}
for atom $n$ in state $a$.
\todo{Please check for consistency.}
\begin{align}
\alpha_{1 1}^{e n} ( x', x )
&=
- \bigl(
\langle e |
\psi_e^{n\dagger} ( x' )
\psi_e^n ( x )
| e \rangle
g^n_{1 1}( x', x )
+
x' \leftrightarrow x
\bigr)
\\
&=
\int \frac{d \omega}{2 \pi} \,
e^{-i \omega (t' - t)} \,
\biggl(
\frac {-1} {\omega^n + \omega + i\, 0^+}
+
\frac {-1} {\omega^n - \omega + i\, 0^+}
\biggr)
\end{align}
The Feynman polarizability
$
\alpha_{1 1}^{e n}(\omega)
$
of the excited atom has a pole in the upper right quadrant of
the $\omega$-plane which leads to a residue (``resonant contribution'')
when integration
contours are shifted to the positive imaginary axis.
The non-equilibrium polarizability gives only a resonant contribution
\chdid{minus sign}
\begin{align}
\alpha_{1 2}^{e n}( x', x )
&=
i \langle e |
\psi_e^{n\dagger} ( x )
\psi_g^n( x )
\psi_e^n ( x' )
\psi_g^{n\dagger} ( x' )
| e \rangle
&=
\int \frac {d \omega}{2 \pi}\,
e^{-i \omega (t' - t)}\,
2 \pi i \,
\delta (\omega - \omega^n)
\;.
\label{eqn:pi_12}
\end{align}
To illustrate the link to the retarded polarizability, we form the combination
\chdid{note the minus sign, conforming with the relation for $D$}
\begin{align}
\alpha^{en}_{1 1} ( x', x ) - \alpha^{en}_{1 2} ( x', x )
&=
-\int \frac {d \omega}{2 \pi}
e^{-i \omega (t' - t)}
\biggl(
\frac{1}{\omega^n + \omega +i\, 0^+}
+
\frac{1}{\omega^n - \omega - i\, 0^+}
\biggr)
\nonumbe
\\& =
\int \frac {d \omega}{2 \pi}
e^{-i \omega (t' - t)}
\alpha^{e}_{\rm R} (\omega, \V{r}_1, \V{r}_2)
\;,
\label{eqn:p11_def_b}
\end{align}
which has poles in the lower half-plane only, as it should.
For ground-state atoms, replace $\omega^n \leftrightarrow -\omega^n$
in these expressions, cf. \pr{eqn:alpha11} for $\alpha^{gn}_{1 1}$.
\section{Green functions of the free electromagnetic field}
\label{sec:free_photons}
We use the following Keldysh--Green functions for the photon field.
To illustrate the formalism, we allow in this appendix for a thermal state
with inverse temperature $\beta$ whence
\chdid{$x',x$ notation; sign in $\bar n( \pm \omega )$ fixed to conform
with (anti)normal ordered results} \todo{HH: \Large$\checkmark$}
\begin{alignat*}{4}
D_{1 1} (x', x)
&=
i\langle
T\bigl\{
E( x' ) E( x )
\bigr\}
\rangle_{\beta}
&\, =& ~~~~~\int\frac{d\omega}{2 \pi} e^{-i \omega( t' - t )}
D_{1 1} (\omega, \V r', \V r )
\;,
\\
D_{1 2} (x', x)
&=
i\langle
E( x ) E( x' )
\rangle_{\beta}
&=& ~~~ i \int\frac{d\omega}{2 \pi} e^{-i \omega( t' - t )}
2 \bar{n}( \omega )\,\mbox{Im}[D_{R} (\omega, \V r', \V r )]
\;,
\\
D_{2 1} (x', x)
&=
i\langle
E( x' ) E( x )
\rangle_{\beta}
&=& -i \int\frac{d\omega}{2 \pi} e^{-i \omega( t' - t )}
2 \bar{n}(-\omega)\,\mbox{Im}[D_{R} (\omega, \V r', \V r )]
\;.
\end{alignat*}
Note that the thermal occupation number becomes
$n(\pm \omega) \to \mp \theta(\mp \omega)$
\chdid{factor $2$ removed} \todo{HH: \Large$\checkmark$}
at zero temperature. The Fourier transforms $D_{1 2}(\omega)$
and $D_{2 1}(\omega)$ are then only supported by
negative / positive frequencies, respectively. In the rest of the paper
we always use this limit. The general case given above follows from
the fluctuation-dissipation theorem~\cite{vanLeeuwen_2006,Agarwal_1975a}.
The link to the retarded and advanced Green functions is provided by
the relations %
$D_{\rm R} = D_{1 1} - D_{1 2}$ and
$D_{\rm A} = D_{1 1} - D_{2 1}.
$
The Feynman propagator in the frequency domain is, therefore,
\begin{align}
D_{1 1}(\omega, \V r', \V r )
=
\SimplePhoton{-0.1ex}{0.07\textwidth}
= \mbox{Re} [D_R(\omega, \V r', \V r ] + i \coth\left(\beta \omega/2\right)\mbox{Im} [D_R(\omega, \V r', \V r )] ~.
\label{eq:photon-11-finite-T}
\end{align}
The retarded Green tensor in a linear and isotropic medium
depends only on the difference $\V r' - \V r$ and is given by
\begin{align}
D_{R}(\omega, \V r, \V 0 ) = & \frac{4 \pi}{3} \delta(r) \mathbbm{1} + \frac{\omega^2}{c^2}\frac{ e^{i k r}}{r}
\left[\left(1 + \frac{i}{k r} - \frac{1}{k^2 r^2}\right)\mathbbm{1}
- \left(1 + \frac{3i}{k r} - \frac{3}{k^2 r^2}\right)
\hat{\V r} \otimes \hat{\V r} \right]~,
\nonumber
\end{align}
where $\hat{ \V r } = \V r / |r|$, and
$k = \sqrt{\varepsilon(\omega) \mu(\omega)} \omega / c$
(${\rm Im}[k] > 0$) is the wave vector in the medium.
Obviously, $[ D_R(\omega, \V r, \V 0) ]^2$ oscillates at half
the medium wavelength, while $|D_R(\omega, \V r, \V 0)|^2$ does not.
In the paper, $D_R( \omega, \V r_B, \V r_A )$ is evaluated by contracting
the tensor above from left and right with the transition dipoles
${\bf d}^B$ and ${\bf d}^A$.
\vspace{-.5ex}
\end{appendix}
{\footnotesize
|
\section{Introduction}
The Astrophysics Data System (ADS) is the premier literature resource for the astronomical
community. It maintains three bibliographic databases containing
roughly 9 million records, 4.5 million scanned pages, and 1.2 million
fulltext articles. Integrated in its databases, the ADS provides
access and pointers to a wealth of external resources, including
electronic articles, data catalogs and archives. While virtually all active
astronomers use the Astrophysics \textit{Data} System as a literature resource, very few know that
it was originally intended as a
\textit{data} resource \citep{ads_orig}.
Nowadays, the ADS is not a data repository \textit{per se}, yet it implicitly contains valuable holdings of astronomical
data, in the form of images, tables and object references contained
within articles. The objective of the ADSASS effort is to
extract these data and make them discoverable and available
through existing data viewers. The outcome of the ADSASS project consists of
a data layer, usable by any astronomical program, which documents,
links, and aggregates information about several astronomical data
pointers.
We focus on the extraction of three categories of data. First, we plan
to assemble astronomical object references provided and curated by
external databases and ``astrotag'' every article in the ADS. Similar
to \textit{geotags}---which are used to reference earth-based locations---
\textit{astrotags} are spatial and temporal annotations about celestial objects.
Second, we employ optical data, in the form of images extracted
from ADS articles to ``astroreference''
every image in the ADS fulltext corpus. Similar to \textit{georeferencing}---which refers to
the alignment of a map for overlay within a given earth-based
coordinate system--- \textit{astroreferencing} allows astrometric
alignment of an image of the sky according to given coordinates, orientation, and pixel scale.
Third, we inspect the metadata of non-optical images (such as text and image caption) to assign
programmatically a position or a source name.
These data sources are then assembled to generate the ADS All-Sky
Survey layer (ADSASS) composed of an all-sky literature heatmap, and an
historical data layer. In this article, we present the technical implementation and proposed
use of the ADSASS.
\section{Technical implementation}
\begin{figure}[h!]
\centering
\includegraphics[width=1.0\textwidth]{adsass}
\caption{The ADS All-Sky Survey: Technical implementation}
\label{fig:techplan}
\end{figure}
The procedures, data sources, and outcomes that are part of the
ADS All-Sky Survey project are depicted, in a summarized
format, in Figure \ref{fig:techplan}. The principal data sources are the
NASA ADS bibliographic database, and the SIMBAD
object catalogue, maintained by the Centre de Données astronomiques de
Strasbourg (CDS).
The first step of the astrotagging procedure involves using all the
holdings of the ADS library for which object references exist to populate a
database of astrotagged literature. This mechanism of article-object
matching generates an initial database that associates ADS articles with their manually curated
object references provided by CDS. The purpose of this first step is to aggregate and publish the
conceptual linkages that already exist between ADS and CDS
holdings. The outcome of this article-object matching procedure is a database of
astrotagged literature. For example, in this procedure a paper by astronomer John
Huchra which discusses the \texttt{M31} \citep{1987AJ.....93..779H} would be
astrotagged with the following elements: a) the referenced objects, each identified by their SIMBAD reference name
(e.g., M 31), and co-ordinates (e.g., 00 42 44.330, +41 16
07.50), and b) a timestamp, identified as the year of publication of the
article (e.g., 1987). However, this article-object matching procedure is only possible for
the portion of the ADS database for which an object
categorization exists. In steps 2 and 3, we
attempt to annotate those papers/images for which no classification is available by automatically inspecting articles for object and/or
coordinate information.
As shown in Figure \ref{fig:techplan}, we plan to extract images from all fulltext
articles in the ADS. For the
purpose of this project, extracted images of interest fall in two
categories: \textit{optical images} and \textit{non-optical images}.
\begin{figure}[h!]
\centering
\includegraphics[width=1.0\textwidth]{opticalnonoptical}
\caption{(a) On the left: a sample optical image from Indriolo et
al. (2010) --- (b): On the right: a sample non-optical image from O' Sullivan et al. (2011)}
\label{opticalnonoptical}
\end{figure}
A sample extracted optical image is presented in the left portion of
Figure \ref{opticalnonoptical} (a). The image
was extracted from a recent article by
\cite{2010ApJ...724.1357I}. When fed to \url{astrometry.net}
\citep{lang2010}, this image is correctly resolved as:
\begin{footnotesize}
\begin{quote}
RA,Dec center: (06:17:4.161, +22:33:23.918); Orientation:
-179.77 deg
E of N; Pixel scale: 4.56 arcsec/pixel; Field size: 77.89 x 58.19
arcminutes; Field contains: Propus (ηGem), IC 443.
\end{quote}
\end{footnotesize}
This reference information (supernova remnant IC443 and the triple-star
system Propus=ηGem) would become part of the astrotags attached to this article,
in addition to a time stamp given by publication date. In this case,
however, in addition to astrotagging the article, we would also
astroreference the image, using the coordinates, orientation, and
pixel scale returned by \url{astrometry.net}. In this way, a database of
literature-extracted astroreferenced images is populated (step 2 of Figure \ref{fig:techplan}).
When astrometric measurement fails, enough textual metadata may be
available to assign either a position or a source name, and optionally
a platescale and waveband to the images. An example of non-optical image is a metal abundance map from a
recent paper by \cite{2011MNRAS.411.1833O}, shown in Figure
\ref{opticalnonoptical} (b). This image cannot be calibrated
astrometrically, yet, the figure caption gives enough information
to extract the needed source name and wavelength metadata. In this
case, it is clear from the caption that the source depicted in this
cluster is the AWM 4 cluster of galaxies. SIMBAD or NED can tell us
that AWM4 is also known as RXC J1604.9+23559, with coordinates RA: 16
04 57.0, DEC: +23 55 14. Thus, using figure caption metadata, this
article/figure would be astrotagged with name and coordinate
information relative to the aforementioned catalog objects (step 3 of Figure \ref{fig:techplan}).
\section{The ADS All-Sky Survey in action}
The ADSASS will yield two products: an \textit{all-sky literature
heatmap}, built from the astrotagged literature database, showing what parts of the sky have been
written about in the literature, and in what contexts; and an
\textit{historical data layer}, built from the astroreferenced image database, offering literature-extracted images for
analysis and overlay on contextual images. Thanks to the proliferation of virtual observatory data
viewing systems over the past few years, the options for visualizing
these layers are many, and they include compatible all-sky viewers
such as the WorldWide Telescope (Microsoft), Aladin (CDS), Google Sky (Google), and MASTview10 (NASA).
This composite ADSASS layer promises to enable and enhance astronomical research
in many ways. First, the ADSASS will draw on the constantly-updated full inventory of
astrotagged ADS literature, to enable on-demand faceted \textit{visual knowledge
discovery} for any target or
topic. For example, a graduate student will be able to generate on the fly a
heatmap of the Orion molecular cloud to figure out which parts of the cloud have existing
molecular line data, where they are missing, and where specific
science topics (e.g., jets and outflows) have been explored in
published works.
Second, the faceted heat map of the ADSASS will enable new forms of
\textit{data discovery}. Consider the problem of discovering published data on ``the radial
velocities of young stars'' in a particular region of interest. A
purely text-based literature search using those words would miss
targets of interest only mentioned in the literature as, for example, ``host stars'' in
``exoplanet surveys''. Instead of this text-only search, exploring the faceted
heat map of the ADSASS, researchers will be able to traverse related
data and objects using a visualization-aided literature search.
Third, researchers’ investigations of the ADSASS historical data layer will
facilitate the \textit{study of astrophysical events}, thanks to its synoptic image view.
Eruptive events, such as an accretion driven outburst in a very young star, reveal objects
that were previously invisible. Synoptic monitoring surveys such as
the Palomar Transient Factory are designed to identify these eruptive
events but on relatively short timescales. The historical data layer will capture and expose data that was
never saved as such in the first place, extending the time baseline
for many parts of the sky back almost 100 years (much longer than the
epoch provided by the Digital Sky Surveys), and providing a framework
to identify solitary or recurrent astrophysical events.
\section{Conclusion}
The ADS All-Sky Survey reunites data and literature holdings
into a powerful astronomical research tool. The mission of the ADSASS
is to ``liberate'' data sources that are contained in scientific papers
as images, tables, object references, and metadata. The
publication of these data in the form of a faceted heat map of the sky
and a historical data layer will enable new forms of astronomical
research and ultimately augment the intrinsic value of the scientific
papers in which those data were embedded.
\acknowledgements The ADS All-Sky Survey was selected for funding as
part of the NASA Research Opportunities in Space and Earth Sciences
(ROSES-2011) for the Astrophysics Data Analysis Program (ADAP). The
authors would like to thank collaborators and project supporters
T. Boch of CDS, A. Accomazzi of ADS, D. Hogg of NYU,
J. Fay of Microsoft Research, and A. Conti of STScI.
\bibliographystyle{asp2010}
|
\section{Introduction}
\label{s1}
An element of $\ensuremath{\mathbb{Z}_n}$ is a \emph{unit} of $\ensuremath{\mathbb{Z}_n}$ if $x$ and $n$ are
co-prime. If the prime-power decomposition of $n$ is
$n = p^\alpha q^\beta r^\gamma \cdots$ where $p, q, r, \ldots\ $ are
distinct primes, then the number of units is given by Euler's
\emph{totient function} $\phi_n = |\ensuremath{\mathbb{U}_n}| =$
$(p-1)p^{\alpha - 1} \cdot (q-1)q^{\beta - 1}
\cdot (r-1)r^{\gamma - 1} \cdots\,$
where $\ensuremath{\mathbb{U}_n}$ denotes the group of units \cite[Chap.~5]{JJ}
We recall the structure of $\ensuremath{\mathbb{U}_n}$:
\begin{itemize}
\item If $n=p^\alpha q^\beta r^\gamma\cdots$, where $p,q,r,\ldots$ are primes,
then
\[\ensuremath{\mathbb{U}_n}=\mathbb{U}_{p^\alpha}\times \mathbb{U}_{q^\beta}\times
\mathbb{U}_{r^\gamma}\times\cdots\ ;\]
\item If $p$ is an odd prime, then $\mathbb{U}_{p^\alpha}$ is cyclic of
order $\phi_{p^\alpha}=p^{\alpha-1}(p-1)$;
\item $\mathbb{U}_{2^\alpha}$ is cyclic of order $2^{\alpha-1}$ if
$\alpha\le2$, and is isomorphic to $C_2\times C_{2^{\alpha-2}}$ otherwise.
\end{itemize}
In particular, although $\ensuremath{\mathbb{U}_n}$ may be expressible as a direct product of
cyclic groups in many different ways, the smallest number of factors is
equal to the number $k$ of prime divisors of $n$ if $n$ is odd, or $k-1$
if $n$ is twice odd, or $k$ if $n$ is four times odd, or $k+1$ otherwise.
The largest number of factors depends on the prime decompositions of $p-1$
for the prime divisors $p$ of $n$. In particular, if $n$ is prime, then
the maximum number of factors is the number of distinct prime divisors of
$n-1$.
We are mainly concerned here with cases where $n$ is odd and $\ensuremath{\mathbb{U}_n}$ is the
product of three cyclic factors. As noted, this forces strong conditions on
$n$: it should have at most three prime divisors; and if $n$ is prime, then
$n-1$ should have at least three prime divisors.
If we can write
$\ensuremath{\mathbb{U}_n} = \langle x \rangle \times \langle y \rangle \times \langle z \rangle$
where $\langle x \rangle$ denotes the subgroup generated by $x$ and the
symbol $\,\times\,$ connotes a direct product, then we have a
\emph{three-factor decomposition} of $\ensuremath{\mathbb{U}_n}$. If the orders \mod{n}
of $x$, $y$ and $z$ are respectively $a$,~$b$~and~$c$, and we need to
specify them succinctly, we use the notation
\[ \ensuremath{\mathbb{U}_n} = \cy{x}{a}\times\cy{y}{b}\times\cy{z}{c} \]
and $\ensuremath{\mathrm{ord}}_n(x) = a$ \textit{etc.}
Irrespective of whether $n$ is prime, a prime power with exponent ${}> 1$,
or a composite, we can sometimes write
\begin{eqnarray}
\ensuremath{\mathbb{U}_n} = \langle x \rangle \times \langle x+k \rangle
\times \langle x+2k \rangle
\label{sct1}
\end{eqnarray}
for some $x,k\in\ensuremath{\mathbb{Z}_n}$, so that the generators are in
arithmetic progression (AP). We then have a \emph{three-factor} AP
\emph{decomposition} of $\ensuremath{\mathbb{U}_n}$, which we abbreviate to a
``3AP decomposition'' of $\ensuremath{\mathbb{U}_n}$. Notable examples~are
\[ \uu{61} = \cy{9}{5} \times \cy{11}{4} \times \cy{13}{3}\ \, \]
and
\[ \uu{911} = \cy{196}{13} \times \cy{550}{10} \times \cy{904}{7}\ , \]
where the orders, like the generators, are in arithmetic progression;
\[ \uu{455} = \cy{92}{4} \times \cy{93}{12} \times \cy{94}{6}\ , \]
where the generators are consecutive integers and the orders are all even,
each being one less than a prime factor of $455$;
\[ \uu{91} = \cy{9}{3} \times \cy{18}{12} \times \cy{27}{2}
= \cy{87}{6} \times \cy{83}{4} \times \cy{79}{3} \]
and
\[ \uu{65} = \cy{61}{3} \times \cy{57}{4} \times \cy{53}{4}\ , \]
where $x = k$ in each of the 3AP decompositions; and
\[ \uu{703} = \cy{700}{9} \times \cy{701}{36} \times \cy{702}{2}\ , \]
where the generators are respectively ${-3}$, ${-2}$ and ${-1}$
\mod{703 = 19 \times 37}.
Our specification (1) may be realised even if the values $x$, $x+k$ and
$x+2k$, when reduced \mod{n} to lie in the interval $[1,n-1]$, are
not in arithmetic progression in $\ensuremath{\mathbb{Z}}$. Thus we have
\[ \uu{31} = \cy{30}{2} \times \cy{2}{5} \times \cy{5}{3} \]
with $k \equiv 3$ \mod{31}. We could equally have written this as
\[ \uu{31} = \cy{5}{3} \times \cy{2}{5} \times \cy{30}{2} \]
with $k \equiv 28 \equiv {-3}$ \mod{31}. Faced with such a choice
between two equivalent representations, we leave ourselves free to choose
whichever seems the more convenient in the context in which it arises.
A rule to make the choice with $0 < k < (n-1)/2$ would be unsatisfactory,
especially as some decompositions~(\ref{sct1}) fall into infinite series
within which $k$ lies variously in $(0,(n-1)/2)$ and in $((n-1)/2,n)$.
Examples where the two \textbf{outer} generators differ by $1$ include
\[ \uu{275} = \cy{136}{5} \times \cy{-1}{2} \times \cy{137}{20} \]
and
\[ \uu{775} = \cy{386}{15} \times \cy{-1}{2} \times \cy{387}{20}\ . \]
We now note three possibilities:
\begin{description}
\item{[A]}
For some values of $n$, different 3AP decompositions of $\ensuremath{\mathbb{U}_n}$ may
arise for different factorisations $\phi_n = a \cdot b \cdot c$. Thus we have
\begin{eqnarray*}
\uu{211} & = &
\cy{15}{6} \times \cy{107}{5} \times \cy{199}{7} \\
& = &
\cy{58}{7} \times \cy{134}{15} \times \cy{210}{2} \\
& = &
\cy{196}{3} \times \cy{203}{35}\times\cy{210}{2} \ .
\end{eqnarray*}
\item{[B]}
For a fixed factorisation $\phi_n = a \cdot b \cdot c$ for a fixed $n$,
we may have different 3AP decompositions
$\langle x \rangle \times \langle y \rangle \times \langle z \rangle$
of $\ensuremath{\mathbb{U}_n}$ where the values of $\ensuremath{\mathrm{ord}}_n(y)$ are different members of
$\{a, b, c\}$. Thus we have
\begin{eqnarray*}
\uu{31} & = &
\cy{30}{2} \times \cy{2}{5} \times \cy{5}{3} \\
& = &
\cy{25}{3} \times \cy{30}{2} \times \cy{4}{5}
\end{eqnarray*}
and
\begin{eqnarray*}
\uu{547} & = &
\cy{40}{3} \times \cy{172}{26} \times \cy{304}{7} \\
& = &
\cy{40}{3} \times \cy{544}{7} \times \cy{501}{26} \\
& = &
\cy{520}{7} \times \cy{40}{3} \times \cy{107}{26}\ .
\end{eqnarray*}
\item{[C]}
We may have different 3AP decompositions of $\ensuremath{\mathbb{U}_n}$ for a fixed
ordering of the terms of a fixed factorisation $\phi_n = a \cdot b \cdot c$
for a fixed~$n$. Thus we~have
\begin{eqnarray*}
\uu{191} & = &
\cy{39}{5} \times \cy{190}{2} \times \cy{150}{19} \\
& = &
\cy{184}{5} \times \cy{190}{2} \times \cy{5}{19}
\end{eqnarray*}
where $184 \equiv 39^2$ \mod{191} and $5 \equiv 150^{-2}$ \mod{191}.
\end{description}
\paragraph{Problem 1} Find a series of primes that behave like~$191$,
with $2$ as the order of the middle generator. (Contenders for inclusion
in the series are $n = 191,\ 271$ and $523$. A possible series with
the order $2$ for an outer generator might cover $n = 331,\ 379,\ 443$
and $647$.)
\medskip
Clearly, a 3AP decomposition of $\ensuremath{\mathbb{U}_n}$, where $n$ is prime and
$n > 4$, cannot exist if the
prime-power decomposition of~$n-1$ contains fewer than 3 distinct primes.
Sufficient conditions for the existence of 3AP decompositions of $\ensuremath{\mathbb{U}_n}$
seem to be elusive. Thus only computer search has established that,
in the range $n < 300$, a 3AP decomposition of $\ensuremath{\mathbb{U}_n}$ does not exist for
any of the values $n = 71,\ 127,\ 139,\ 223$ and~$277$. (We here
exclude the ``weak'' 3AP decompositions defined in \S\ref{s2} below.)
For any $n$ with $n > 4$, there is a \emph{primitive root} of $n$ (an
element from $\ensuremath{\mathbb{U}_n}$ that generates all members of $\ensuremath{\mathbb{U}_n}$)
if and only if $n$ is an odd prime power or twice
an odd prime power. In general we write $\lambda_n$ for the maximum
order of a member of $\ensuremath{\mathbb{U}_n}$; if $n$ is odd, with prime power decomposition
$n = p^\alpha q^\beta r^\gamma \cdots$, then
\[ \lambda_n = \textup{lcm}((p-1)p^{\alpha - 1},\ (q-1)q^{\beta - 1},
\ (r-1)r^{\gamma - 1},\ \ldots)\ . \]
We write $\xi_n = \phi_n / \lambda_n$;
as shown in \cite[\S6]{cap}, $\xi_n$ is even if greater than~$1$.
\paragraph{Problem 2} Is there an upper bound on the number of $3$AP
decompositions of $\uu{n}$ in terms of $\xi(n)$? Conversely, for a given
value of $\xi(n)=m$, is it always possible to find $n$ with no $3$AP
decompositions?
Empirically we have found a tendency for larger values of $\xi(n)$ to be
associated with larger numbers of decompositions. The table below,
obtained by computer, gives $D$,
the maximum number of 3AP decompositions of $\uu{n}$, where $n\le1000$ and
$\xi(n)$ is prescribed.
\[\begin{array}{c|rrrrrrrrrrrr}
\xi(n) & 1 & 2 & 4 & 6 & 8 & 10 & 12 & 16 & 18 & 20 & 24 & 36\\\hline
\phantom{|^|}D & 10 & 18 & 96 & 182 & 288 & 262 & 496 &
384 & 276 & 204 & 540 & 2088
\end{array}\]
\section{$n$ prime}
\label{s2}
For $n$ prime, the multiplicative group $\ensuremath{\mathbb{U}_n}$ is cyclic, and so if it is
expressed as a direct product, the factors must have pairwise co-prime orders.
\subsection{The case $n-1 = 2 \cdot 3 \cdot m$}
\label{ss21}
We first prove three theorems that apply for prime values $n$ such
that the factors in a 3AP decomposition of $\ensuremath{\mathbb{U}_n}$ have orders
$2$, $3$ and $m$, where $2$, $3$ and $m$ are pairwise co-prime.
The first of these theorems is closely linked to Theorem~2.7 of
\cite{consec}. We begin with some preliminary remarks.
Our assumption on $n$ implies that $n\equiv 7$ or $31$ \mod{36},
and $n>7$. In particular, since $n\equiv3$ \mod{4}, the
quadratic residues have odd order, and the non-residues have even order.
In our Theorems, we will be interested in the solutions of the quadratic
equation $x^2+3x+3=0$ in $\ensuremath{\mathbb{Z}_n}$. Its discriminant is $-3$, which
(by Quadratic Reciprocity \cite[\S7.4]{JJ}) is a square in $\ensuremath{\mathbb{Z}_n}$,
so the quadratic has two roots in~$\ensuremath{\mathbb{Z}_n}$.
The product of the roots is $3$, which is a non-square; so one root
(say~$x_1$) has odd order, and the other (say $x_2$) has even order.
We also note that the values $y_1=x_1+1$ and $y_2=x_2+1$ satisfy the
quadratic equation $y^2+y+1=0$, and so $\ensuremath{\mathrm{ord}}_n(y_1)=\ensuremath{\mathrm{ord}}_n(y_2)=3$.
\begin{theorem}
Let $n$ be a prime satisfying $n \equiv 7$ or $31$ \mod{36},
$n>7$. Suppose that the elements $x_1$ and $x_2$ from $\ensuremath{\mathbb{U}_n}$ that satisfy
$x^2 + 3x + 3 \equiv 0$ \mod{n} are such that
$\ensuremath{\mathrm{ord}}_n(x_1) = (n-1)/6$. Then $\ensuremath{\mathrm{ord}}_n(-(x_1 +2 )) = 3$ and so
\[ \ensuremath{\mathbb{U}_n} = \cy{-x_1 - 2}{3} \times \cy{-1}{2} \times \cy{x_1}{m} \]
where $m = (n-1)/6$.
\label{t21}
\end{theorem}
\begin{pf}
As noted above, $3 = \ensuremath{\mathrm{ord}}_n(x_2 + 1)$, and $x_2+1=-x_1-2$, since
$x_1+x_2=-3$.
\end{pf}
\paragraph{Coverage}
In the range $n < 1000$, Theorem~\ref{t21} covers values as follows:
\renewcommand{\arraystretch}{1.4}
\begin{eqnarray*}
n = 31: && \cy{25}{3} \times \cy{30}{2} \times \cy{4}{5} \\
n = 43: && \cy{6}{3} \times \cy{42}{2} \times \cy{35}{7} \\
n = 79: && \cy{55}{3} \times \cy{78}{2} \times \cy{22}{13} \\
n = 211: && \cy{196}{3} \times \cy{210}{2} \times \cy{13}{35} \\
n = 463: && \cy{21}{3} \times \cy{462}{2} \times \cy{440}{77} \\
n = 571: && \cy{109}{3} \times \cy{570}{2} \times \cy{460}{95}\\
n = 751: && \cy{678}{3} \times \cy{750}{2} \times \cy{71}{125}\\
n = 907: && \cy{522}{3} \times \cy{906}{2} \times \cy{383}{151}
\end{eqnarray*}
\begin{theorem}
Let $n$ be a prime satisfying $n \equiv 7$ or $31$ \mod{36},
$n>7$. Suppose that the elements $x_1$ and $x_2$ from $\ensuremath{\mathbb{U}_n}$ that
satisfy $x^2 + 3x + 3 \equiv 0$ \mod{n} are such that
$\ensuremath{\mathrm{ord}}_n(x_1) = (n-1)/2$ and $\ensuremath{\mathrm{ord}}_n(x_2) = (n-1)$. Then
$\ensuremath{\mathrm{ord}}_n(x_2 + 1) = 3$ and $\ensuremath{\mathrm{ord}}_n(2x_2 + 3) = (n-1)/6$, so that
\[ \ensuremath{\mathbb{U}_n} = \cy{2x_2 + 3}{m} \times \cy{x_2 + 1}{3} \times \cy{-1}{2} \]
where $m = (n-1)/6$.
\label{t22}
\end{theorem}
\begin{pf}
Since $\ensuremath{\mathbb{U}_n}=\langle-1\rangle_2\times\langle x_1+1\rangle\times C_{(n-1)/6}$,
the hypothesis $\ensuremath{\mathrm{ord}}_n(x_2)=n-1$ shows that
$x_2=-(x_1+1)c$ or $-(x_2+1)c$, where $\ensuremath{\mathrm{ord}}_n(c)=(n-1)/6$.
Hence either $-(x_1+1)x_2$ or $=-(x_2+1)x_2$ has order $(n-1)/6$.
Now $x_1+x_2=-3$ and $x_1x_2=3$, so $-(x_1+1)x_2=-3-x_2=x_1$,
which has order $(n-1)/2$, by assumption. So
\[-(x_2+1)x_2=-x_2^2-x_2=2x_2+3\]
has order $(n-1)/6$.
\end{pf}
\paragraph{Coverage}
In the range $n < 1000$, Theorem~2.2 covers values as follows:
\begin{eqnarray*}
n = 67: && \cy{59}{11} \times \cy{29}{3} \times \cy{66}{2} \\
n = 103: && \cy{10}{17} \times \cy{46}{3} \times \cy{102}{2} \\
n = 151: && \cy{86}{25} \times \cy{118}{3} \times \cy{150}{2} \\
n = 367: && \cy{200}{61} \times \cy{283}{3} \times \cy{366}{2} \\
n = 439: && \cy{343}{73} \times \cy{171}{3} \times \cy{438}{2} \\
n = 499: && \cy{279}{83} \times \cy{139}{3} \times \cy{498}{2} \\
n = 619: && \cy{505}{103} \times \cy{252}{3} \times \cy{618}{2} \\
n = 643: && \cy{355}{107} \times \cy{177}{3} \times \cy{642}{2} \\
n = 727: && \cy{563}{121} \times \cy{281}{3} \times \cy{726}{2} \\
n = 787: && \cy{28}{131} \times \cy{407}{3} \times \cy{786}{2} \\
n = 967: && \cy{682}{162} \times \cy{824}{3} \times \cy{966}{2} \\
\end{eqnarray*}
\begin{theorem}
Let $n$ be a prime satisfying $n \equiv 7$ or $31$ \mod{36}.
Suppose that~$z$ is one of the elements $x_1$ and $x_2$ that satisfy
$x^2 + 3x + 3 \equiv 0$ \mod{n} and that
$\ensuremath{\mathrm{ord}}_n(2^{-1} z) = (n-1)/6$. Then
\[ \ensuremath{\mathbb{U}_n} = \cy{z+1}{3} \times \cy{2^{-1}z}{m} \times \cy{-1}{2} \]
where $m = (n-1)/6$.
\label{t23}
\end{theorem}
\begin{pf} As for Theorem~2.1. But it depends on $n$
whether $z$ is the solution of $x^2 + 3x + 3 \equiv 0$ that has the
larger or smaller order, and whether $z$ is $x_1$ or $x_2$.
\end{pf}
\paragraph{Coverage}
In the range $n < 1000$, Theorem~2.3 covers values as follows:
\begin{center}
$
\begin{array}{rcl}
n = 31: & \cy{5}{3} \times \cy{2}{5} \times \cy{30}{2} & (z = x_1) \\
n = 67: & \cy{29}{3} \times \cy{14}{11} \times \cy{66}{2} & (z = x_2) \\
n = 103: & \cy{56}{3} \times \cy{79}{17} \times \cy{102}{2} & (z = x_1) \\
n = 151: & \cy{32}{3} \times \cy{91}{25} \times \cy{150}{2} & (z = x_1) \\
n = 211: & \cy{196}{3} \times \cy{203}{35} \times \cy{210}{2} & (z = x_2) \\
n = 283: & \cy{44}{3} \times \cy{163}{47} \times \cy{282}{2} & (z = x_2) \\
n = 691: & \cy{437}{3} \times \cy{218}{115} \times \cy{690}{2}& (z = x_2) \\
n = 787: & \cy{407}{3} \times \cy{203}{131} \times \cy{786}{2}& (z = x_2) \\
n = 823: & \cy{648}{3} \times \cy{735}{137} \times \cy{822}{2}& (z = x_1) \\
n = 907: & \cy{522}{3} \times \cy{714}{151} \times \cy{906}{2}& (z = x_2) \\
\end{array}
$
\end{center}
\paragraph{Note 2.1} In the range $n < 1000$,
Theorems~\ref{t21}--\ref{t23} exclude $n = 139$, 223, 331, 547, 607 and 859.
All but one of these has $x$-values $x_1$ and $x_2$ with
$\ensuremath{\mathrm{ord}}_n(x_1) = (n-1)/2$ and $\ensuremath{\mathrm{ord}}_n(x_2) = (n-1)/3$; the exception is
$n = 547$, which has $x_1 = 505$ and $x_2 = 39$, with
$\ensuremath{\mathrm{ord}}_n(x_1) = (n-1)/26$ and $\ensuremath{\mathrm{ord}}_n(x_2) = (n-1)/13$.
\paragraph{Problem 3} It is natural to wonder whether there are infinitely
many primes for which the conditions of one of the above theorems are
satisfied. Here are some thoughts on this. In all cases we seek primes
congruent to $7$ or $31$ \mod{36}; Dirichlet's Theorem \cite[Theorem~2.10]{JJ}
guarantees that infinitely many such primes exist, and indeed they have
density $1/6$ among all primes.
Consider Theorem~\ref{t21}. We require that an element of order $(n-1)/6$
(necess\-arily a sixth power) should satisfy $x^2+3x+3=0$, so there should
be a solution $y$ of the equation $y^{12}+3y^6+3=0$. The Chebotarev density
theorem \cite{Chebotarev}, \cite[section 1.2.2]{Chebotarev2}
guarantees that this equation will
have a solution in an infinite set (indeed, a set of positive density) of
primes. This theorem can further guarantee a set of positive density for
which $x_1$ has six distinct sixth roots (so that $n\equiv1$ \mod{6}) and
$x_2$ is a non-square (so that $n\equiv3$ \mod{4}), but we do not know how
to exclude $n\equiv19$ \mod{36}.
A more serious difficulty is that the fact that $x$ is a sixth power
guarantees only that its order divides $(n-1)/6$; it does not seem
easy to show that the order is precisely this value. Clearly this would
be the case if $n=6q+1$ with $q$ prime; but it is not even known whether
infinitely many primes of this form occur.
Of the $1614$ primes less than $10^5$ which are congruent to $7$ or $31$
\mod{36}, there are $494$, $476$ and~$476$ that satisfy the conditions of
Theorems~\ref{t21}--\ref{t23} respectively
\medbreak
We conclude this subsection with a converse to the preceding theorems.
\begin{theorem}
Any \textup{3AP} decomposition of $\ensuremath{\mathbb{U}_n}$ for $n$ prime, in which the
generators have orders $2$, $3$ and $(n-1)/6$, arises as in one of
the three preceding theorems.
\label{t24}
\end{theorem}
\begin{pf}
We already saw that $n$ must be congruent to $7$ or $31$ \mod{36}. The
only element of order~$2$ is $-1$, and the only elements of order~$3$ are
$x_1+1$ and $x_2+1$, where $x_1$ and $x_2$ are the roots of $x^2+3x+3=0$
(with the convention that $x_1$ has odd order and $x_2$ even order).
The only possibilities for the third generator are thus
$-x_i-3$, $2x_i+3$, or $2^{-1}x_i$, for $i=1$ or $i=2$. We treat the three
cases in turn.
In the first case, since $x_1+x_2=-3$, we have $-x_1-3=x_2$ and
\emph{vice versa}. Since $x_2$ has even order by our convention, we must
have $i=2$, and the generators are $x_1$, $-1$ and $x_2+1$; the requirement
is that $x_1$ has order $(n-1)/6$.
In the second case, we assume that $2x_i+3$ has order $(n-1)/6$, and have to
prove that $i=2$ and that the orders of $x_1$ and $x_2$ are $(n-1)/2$ and
$(n-1)$ respectively. Let $j=3-i$. From the proof of Theorem~\ref{t22}, we see
that $2x_i+3=-(x_i+1)x_i$, so that $x_i=-(x_j+1)(2x_i+3)$, the product of
elements of orders $2$, $3$ and $(n-1)/6$; so $x_i$ has order $n-1$. Thus
$i=2$. Now $(x_1+1)x_1=2x_2+3$ has order $(n-1)/6$, so $x_1=(x_2+1)(2x_2+3)$
has order $(n-1)/2$.
Finally, the third case obviously gives the situation of Theorem~\ref{t23}.
\end{pf}
\subsection{The case $n-1 = 3 \cdot 4 \cdot \mu$}
\label{ss22}
We now prove two theorems that apply for prime values $n$ such
that the factors in a 3AP decomposition of $\ensuremath{\mathbb{U}_n}$ have orders
$3$, $4$ and $\mu$ where $3$, $4$ and $\mu$ are pairwise co-prime.
We give no theorem for the situation where $\mu$ is the order of the
middle generator. This case can occur; the smallest example is for $n=997$.
\begin{theorem}
Let $n$ be a prime satisfying $n \equiv 13$, $61$, $85$ or $133$
\mod{144}, $n > 13$. Suppose that there is an element $x$ from
$\ensuremath{\mathbb{U}_n}$ such that $x^2 + 3x + 3 \equiv 0$ \mod{n} and such
that there is also an element $k$ with $\ensuremath{\mathrm{ord}}_n(x+1+k) = 4$ and
$\ensuremath{\mathrm{ord}}_n(x+1+2k) = (n-1)/12$. Then
\[ \ensuremath{\mathbb{U}_n} = \cy{x+1}{3} \times \cy{x+1+k}{4} \times \cy{x+1 +2k}{\mu} \]
where $\mu = (n-1)/12$.
\label{t25}
\end{theorem}
\begin{pf}
As for Theorem~2.1. Note that the condition on $x+1+k$
can be written $(x+1+k)^2\equiv-1$ \mod{n}.
\end{pf}
\paragraph{Coverage}
In the range $n < 1000$, Theorem~\ref{t25} covers values as follows:
\begin{center}
$
\begin{array}{rcl}
n & \textup{3AP decomposition of }\ensuremath{\mathbb{U}_n} & \ensuremath{\mathrm{ord}}_n(x) \\
\hline
61 & \langle 13 \rangle_3 \times \langle 11 \rangle_4
\times \langle 9 \rangle_5 & 15 = (n-1)/4 \\
349 & \langle 122 \rangle_3 \times \langle 213 \rangle_4
\times \langle 304 \rangle_{29} & 58 = (n-1)/6 \\
661 & \left\{ \begin{array}{@{}l@{}}
\langle 364 \rangle_3 \times \langle 106 \rangle_4
\times \langle 509 \rangle_{55} \\
\langle 364 \rangle_3 \times \langle 555 \rangle_4
\times \langle 85 \rangle_{55}
\end{array} \right\} & 66 = (n-1)/10 \\
\hline
\end{array}
$
\end{center}
For the two examples for $n = 661$, the generators of order 55 are related
by the congruence $85 \equiv 509^3$ \mod{661}.
\begin{theorem}
Let $n$ be a prime satisfying $n \equiv 13$, $61$, $85$ or $133$
\mod{144}, $n > 13$. Suppose that there is an element $x$ from
$\ensuremath{\mathbb{U}_n}$ such that $x^2 + 3x + 3 \equiv 0$ \mod{n} and such
that there is also an element $k$ with $\ensuremath{\mathrm{ord}}_n(x+1+k) = 4$ and
$\ensuremath{\mathrm{ord}}_n(x+1-k) = (n-1)/12$. Then
\[ \ensuremath{\mathbb{U}_n} = \langle x+1-k \rangle_\mu \times \langle x+1 \rangle_3
\times \langle x+1+k \rangle_4 \]
where $\mu = (n-1)/12$.
\label{t26}
\end{theorem}
\begin{pf} As for Theorem~2.1.
\end{pf}
\paragraph{Coverage}
In the range $n < 1000$, Theorem~\ref{t26} covers values as follows:
\begin{center}
$
\begin{array}{rcl}
n & \textup{3AP decomposition of }\ensuremath{\mathbb{U}_n} & \ensuremath{\mathrm{ord}}_n(x) \\
\hline
157 & \langle 153 \rangle_{13} \times \langle 12 \rangle_3
\times \langle 28 \rangle_4 & 39 = (n-1)/4 \\
229 & \langle 161 \rangle_{19} \times \langle 134 \rangle_3
\times \langle 107 \rangle_4 & 228 = (n-1) \\
349 & \langle 31 \rangle_{29} \times \langle 122 \rangle_3
\times \langle 213 \rangle_4 & 58 = (n-1)/6 \\
373 & \langle 91 \rangle_{31} \times \langle 284 \rangle_3
\times \langle 104 \rangle_4 & 93 = (n-1)/4 \\
997 & \langle 226 \rangle_{83} \times \langle 692 \rangle_3
\times \langle 161 \rangle_4 & 498 = (n-1)/2 \\
\hline
\end{array}
$
\end{center}
\medskip
In the range $n < 1000$, Theorems \ref{t25} and \ref{t26} fail to provide
3AP decompositions of $\ensuremath{\mathbb{U}_n}$ for $n = 277$, 421, 709, 733, 853 and 877.
However, 3AP decompositions for $n = 421$ exist for other partitions of $n-1$.
\subsection{The case $n-1 = 2 \cdot 5 \cdot \nu$}
\label{ss23}
We now examine what occurs for primes $n$ such that the factors in a
3AP decomposition of $\ensuremath{\mathbb{U}_n}$ have orders 2, 5 and $\nu$, these orders
being pairwise co-prime.
Amongst primes satisfying $n \equiv 11$, 31, 71 and 91 (mod~100),
$n > 11$, the patterns of occurrence of such 3AP decompositions are
very similar to those reported in~\S\ref{ss21} above.
For each relevant value of $n$ there are 4 elements of order 5; their sum
is ${-1}$ and their product is ${+1}$. Sometimes more than one of
the four can be used.
\paragraph{Type 2.3(a)}
Analogous to the decompositions obtainable via Theorem \ref{t21}, we now
have 3AP decompositions of the form
\[ \ensuremath{\mathbb{U}_n} = \cy{-z-2}{5} \times \cy{-1}{2} \times \cy{z}{\nu}\ . \]
In the range $n < 1000$ they are as follows:
\begin{eqnarray*}
n = 31: && \cy{4}{5} \times \cy{30}{2} \times \cy{25}{3} \\
n = 191: & \left\{ \begin{array}{@{}c@{}} \\ \\ \end{array} \right.
& \begin{array}{@{}l@{}}
\cy{39}{5} \times \cy{190}{2} \times \cy{150}{19} \\
\cy{184}{5} \times \cy{190}{2} \times \cy{5}{19} \\
\end{array} \\
n = 271: & \left\{ \begin{array}{@{}c@{}} \\ \\ \end{array} \right.
& \begin{array}{@{}l@{}}
\cy{10}{5} \times \cy{270}{2} \times \cy{259}{27} \\
\cy{244}{5} \times \cy{270}{2} \times \cy{25}{27} \\
\end{array} \\
n = 431: && \cy{405}{5} \times \cy{430}{2} \times \cy{24}{43} \\
n = 691: && \cy{89}{5} \times \cy{690}{2} \times \cy{600}{69} \\
n = 991: && \cy{799}{5} \times \cy{990}{2} \times \cy{190}{99}
\end{eqnarray*}
\paragraph{Type 2.3(b)}
Analogous to the decompositions obtainable via Theorem \ref{t22}, we
have 3AP decompositions of the form
\[ \ensuremath{\mathbb{U}_n} = \cy{2z+1}{\nu} \times \cy{z}{5} \times \cy{-1}{2}\ . \]
In the range $n < 1000$ they are as follows:
\begin{eqnarray*}
n = 31: && \cy{5}{3} \times \cy{2}{5} \times \cy{30}{2} \\
n = 131: && \cy{107}{13} \times \cy{53}{5} \times \cy{130}{2} \\
n = 311: & \left\{ \begin{array}{@{}c@{}} \\ \\ \end{array} \right.
& \begin{array}{@{}l@{}}
\cy{13}{31} \times \cy{6}{5} \times \cy{310}{2} \\
\cy{105}{31} \times \cy{52}{5} \times \cy{310}{2} \\
\end{array} \\
n = 491: && \cy{203}{49} \times \cy{101}{5} \times \cy{490}{2} \\
n = 811: && \cy{330}{81} \times \cy{570}{5} \times \cy{810}{2} \\
n = 991: && \cy{395}{99} \times \cy{197}{5} \times \cy{990}{2}
\end{eqnarray*}
\paragraph{Type 2.3(c)}
Analogous to the decompositions obtainable via Theorem \ref{t23}, we
have 3AP decompositions of the form
\[ \ensuremath{\mathbb{U}_n} = \cy{2z+1}{5} \times \cy{z}{\nu} \times \cy{-1}{2}\ . \]
In the range $n < 1000$ they are as follows:
\begin{eqnarray*}
n = 271: && \cy{10}{5} \times \cy{140}{27} \times \cy{270}{2} \\
n = 691: && \cy{132}{5} \times \cy{411}{69} \times \cy{690}{2} \\
n = 971: && \cy{803}{5} \times \cy{401}{97} \times \cy{970}{2} \\
n = 991: && \cy{197}{5} \times \cy{98}{99} \times \cy{990}{2}
\end{eqnarray*}
\paragraph{Note 2.2} In the range $n < 1000$, no 3AP decomposition of any
of the types 2.3(a), 2.3(b) or 2.3(c) exists for $n = 71$, $211$, $331$,
$571$, $631$ or $911$.
\subsection{Some double-barrelled cases}
\label{ss24}
Two special cases arise for prime $n$ such that $\ensuremath{\mathbb{U}_n}$ has more than
one 3AP decomposition. These are where we can write either
\[ \ensuremath{\mathbb{U}_n} = \cy{k}{a} \times \cy{k+z}{b} \times \cy{k+2z}{c}
= \cy{k-2z}{b} \times \cy{k}{a} \times \cy{k+2z}{c} \]
or
\[ \ensuremath{\mathbb{U}_n} = \cy{k}{a} \times \cy{k+z}{b} \times \cy{k+2z}{c}
= \cy{k+z}{b} \times \cy{k+2z}{c} \times \cy{k+3z}{a}\ . \]
For the first of these we need $k+z$ and $k - 2z$ to have the same
order \mod{n} and each to be a power of the other \mod{n}.
For the second we need the same relationship between $k$ and $k+3z$.
We have failed to find any theorems to indicate when these cases arise.
In the range $n < 1000$, the occurrences of the first case are these:
\begin{center}
$
\begin{array}{lcccc}
\uu{67} & = & \cy{29}{3} \times \cy{14}{11} \times \cy{66}{2}
& = & \cy{59}{11} \times \cy{29}{3} \times \cy{66}{2} \\
\uu{211} & = & \cy{210}{2} \times \cy{203}{35} \times \cy{196}{3}
& = & \cy{13}{35} \times \cy{210}{2} \times \cy{196}{3} \\
\uu{271} & = & \cy{270}{2} \times \cy{140}{27} \times \cy{10}{5}
& = & \cy{259}{27} \times \cy{270}{2} \times \cy{10}{5} \\
\uu{331} & = & \cy{167}{11} \times \cy{83}{15} \times \cy{330}{2}
& = & \cy{4}{15} \times \cy{167}{11} \times \cy{330}{2} \\
\uu{379} & = & \cy{378}{2} \times \cy{119}{7} \times \cy{239}{27}
& = & \cy{138}{7} \times \cy{378}{2} \times \cy{239}{27} \\
\uu{661} & = & \cy{364}{3} \times \cy{391}{20} \times \cy{418}{11}
& = & \cy{310}{20} \times \cy{364}{3} \times \cy{418}{11} \\
\uu{787} & = & \cy{407}{3} \times \cy{203}{131} \times \cy{786}{2}
& = & \cy{28}{131} \times \cy{407}{3} \times \cy{786}{2} \\
\uu{907} & = & \cy{906}{2} \times \cy{714}{151} \times \cy{522}{3}
& = & \cy{383}{151} \times \cy{906}{2} \times \cy{522}{3} \\
\end{array}
$
\end{center}
whereas the occurrences of the second case are these:
\begin{center}
$
\begin{array}{lcccc}
\uu{349} & = & \cy{31}{29} \times \cy{122}{3} \times \cy{213}{4}
& = & \cy{122}{3} \times \cy{213}{4} \times \cy{304}{29} \\
\uu{599} & = & \cy{578}{23} \times \cy{598}{2} \times \cy{19}{13}
& = & \cy{598}{2} \times \cy{19}{13} \times \cy{39}{23}\ . \\
\end{array}
$
\end{center}
\section {Lifts}
\label{s3}
We now consider how and when an AP decomposition of $\uu{n}$ can be used
to obtain AP decompositions of $\uu{n'}$ where $n'$ is a power of $n$ or
some other multiple of~$n$.
\subsection{Definitions}
\label{ss31}
In this section, we allow \emph{weak} 3AP decompositions
$\uu{n}=\cy{x}{a}\times\cy{y}{b}\times\cy{z}{c}$,
where one of $x,y,z$ is allowed to be $1$ (so that the corresponding
cyclic factor is trivial). For example,
\begin{eqnarray}
\uu{7}=\cy{4}{3}\times\cy{6}{2}\times\cy{1}{1}
\label{sct31}
\end{eqnarray}
and
\[\uu{103}=\cy{1}{1}\times\cy{47}{6}\times\cy{93}{17}\]
are weak 3AP decompositions. Where it aids clarity, we refer to a 3AP
decomposition in the original sense as being \emph{strong}.
If $p$ is a prime satisfying $p \equiv 11$ \mod{12} and $\ensuremath{\mathrm{ord}}_p(3) = (p-1)/2$,
then $\uu{p} = \cy{-1}{2}\times\cy{1}{1}\times\cy{3}{(p-1)/2}$.
Likewise if $p$ is a prime satisfying $p \equiv 7$ \mod{12} and
$\ensuremath{\mathrm{ord}}_p(-3) = (p-1)/2$, then
$\uu{p} = \cy{-3}{(p-1)/2}\times\cy{-1}{2}\times\cy{1}{1}$.
If $n$ divides $n'$, then the map $x\mapsto x$ \mod{n} is a ring epimorphism
from $\zz{n'}$ to $\zz{n}$, and maps $\uu{n'}$ onto $\uu{n}$. It preserves
the property of forming an arithmetic progression. However, it does not
in general map a (weak) 3-AP decomposition of $\uu{n'}$ to a (weak)
3-AP decomposition of $\uu{n}$. (It maps a generating set to a generating
set, but does not necessarily preserve the direct sum decomposition.)
Note in passing that, if $n$ divides $n'$, then $\phi(n)$ divides $\phi(n')$,
the quotient being the order of the kernel of the homomorphism from
$\uu{n'}$ to $\uu{n}$.
Suppose that $n$ divides $n'$, and that the 3AP decompositions
\[\uu{n}=\cy{x}{a}\times\cy{y}{b}\times\cy{z}{c}\]
and
\[\uu{n'}=\cy{x'}{a'}\times\cy{y'}{b'}\times\cy{z'}{c'}\]
satisfy
$x \equiv x'$ \mod{n}, $y' \equiv y$ \mod{n} and $z \equiv z'$ \mod{n}.
Then we call the
second decomposition a \emph{lift} of the first, with \emph{index} $n'/n$.
Note that we must have $a\mid a'$, $b\mid b'$, $c\mid c'$, and
$(a'b'c')/(abc)=\phi(n')/\phi(n)$. For example, the decomposition
\begin{eqnarray}
\uu{49}=\cy{18}{3}\times\cy{48}{2}\times\cy{29}{7}
\label{sct32}
\end{eqnarray}
is a lift of the weak 3AP decomposition (\ref{sct31}) of $\uu{7}$.
We further describe $x'$ as being a \emph{lift} of $x$, and so on.
\subsection{Lifts from $n$ to $np$, with $p$ an odd prime}
\label{ss32}
We are unable to give necessary and sufficient conditions for lifts to
exist. In the remainder of this section, we consider the case where $n'=np$
for an odd prime $p$. (We do not know whether every lift can be obtained
by a sequence of lifts where the indices are primes.)
We subdivide the analysis into three cases.
\paragraph{Case 1: $p^2$ divides $n$.} Let $n=p^km$ ($k \ge 2$) where
$p$ does not divide $m$.
In this case, if an element $x\in\uu{n}$ has $\ensuremath{\mathrm{ord}}_{p^k}(x)$ divisible by $p$,
then any lift $x'$ of $x$ satisfies $\ensuremath{\mathrm{ord}}_{n'}(x')=p\,\ensuremath{\mathrm{ord}}_n(x)$. So at most
one of $x,y,z$ can satisfy this condition if there is a lift which is a
3AP decomposition. Conversely, if, say, $x$~satisfies the condition but
$y$ and $z$ do not, then we can choose lifts $y'$
and $z'$ of $y$ and $z$ satisfying $\ensuremath{\mathrm{ord}}_{n'}(y')=\ensuremath{\mathrm{ord}}_n(y)$ and
$\ensuremath{\mathrm{ord}}_{n'}(z')=\ensuremath{\mathrm{ord}}_n(z)$, and the unique lift $x'$ of $x$ such that
$(x',y',z')$ is an AP in $\zz{n'}$, to obtain a 3AP decomposition
of $\uu{n'}$.
For example, below (Case 2, Subcase 2.2) we find a decomposition
\[\uu{275}=\cy{274}{2} \times \cy{166}{5} \times \cy{58}{20}.\]
This cannot be lifted to a 3AP decomposition of $\uu{n'}$ with
$n' = 5 \times 275$. On the other hand, the decomposition (\ref{sct32})
can be lifted to
\[\uu{343}=\cy{18}{3}\times\cy{342}{2}\times\cy{323}{49}.\]
\paragraph{Case 2:} $p$ divides $n$ but $p^2$ does not. We further subdivide
into four cases according to how many of the orders $a,b,c$ are divisible
by~$p$. Note that $\phi(np)=\phi(n)p$.
For any
$x\in\uu{n}$, with $\ensuremath{\mathrm{ord}}_n(x)=a$, the lifts of $x$ to $\uu{pn}$ belong to
an extension of $C_p$ by $C_a$, which is isomorphic to $C_p\times C_a$.
Hence, if $p$ does not divide $a$, then one of these lifts (which
we call the \emph{special lift}) has order~$a$, and the other $p-1$ have
order~$pa$; while if $p$ divides $a$, then all have order~$a$.
\subparagraph{Subcase 2.1:} None of $a,b,c$ is divisible by~$p$. If a lift
is a 3AP decomposition, then two of $x,y,z$ lift to elements of the same
orders (and so must be special lifts), while the third lifts to an element
with $p$ times the order. Let
$x',y',z'$ be the special lifts of $x,y,z$. We call the decomposition
\emph{unproductive} if $(x',y',z')$ is an AP in $\zz{np}$. In this case,
there is no lift to a 3AP decomposition of $\uu{np}$. In the contrary
\emph{productive} case, there are three lifts which are 3AP decompositions
since we may choose for which two of $x,y,z$ we use special lifts, and the
third lift is determined by the AP requirement.
Thus, if we start from the productive strong 3AP decomposition
\[\uu{31} = \cy{25}{3} \times \cy{30}{2} \times \cy{4}{5}\]
we obtain the three lifts
\begin{eqnarray*}
\uu{31^2}
&=& \cy{521}{3} \times \cy{960}{2} \times \cy{438}{155} \\
&=& \cy{521}{3} \times \cy{526}{62} \times \cy{531}{5} \\
&=& \cy{428}{93} \times \cy{960}{2} \times \cy{531}{5}\,.
\end{eqnarray*}
Likewise we can start from the productive strong 3AP decomposition
\begin{eqnarray*}
\uu{35} & = & \cy{11}{3} \times \cy{34}{2} \times \cy{22}{4}\ .
\end{eqnarray*}
to obtain the lifts
\begin{eqnarray*}
\uu{245} & = & \cy{116}{3} \times \cy{244}{2} \times \cy{127}{28} \\
& = & \cy{116}{3} \times \cy{34}{14} \times \cy{197}{4} \\
& = & \cy{46}{21} \times \cy{244}{2} \times \cy{197}{4}
\end{eqnarray*}
and
\begin{eqnarray*}
\uu{175} & = & \cy{151}{3} \times \cy{174}{2} \times \cy{22}{20} \\
& = & \cy{151}{3} \times \cy{104}{10} \times \cy{57}{4} \\
& = & \cy{116}{15} \times \cy{174}{2} \times \cy{57}{4}\ .
\end{eqnarray*}
If we start from a productive weak 3AP decomposition, then two of
the three lifts are weak but the third (where the identity lifts to an
element of order~$p$) is strong. This happens in the example (\ref{sct32})
of a 3AP decomposition of $\uu{49}$; the two corresponding weak 3AP
decompositions lifted from (\ref{sct31}) are
\[\uu{49}=\cy{18}{3}\times\cy{34}{14}\times\cy{1}{1}
=\cy{46}{21}\times\cy{48}{2}\times\cy{1}{1}.\]
Now consider the prime $n = 379$. The strong 3AP decomposition
\begin{eqnarray}
\uu{379} &=& \cy{239}{27}\times\cy{378}{2}\times\cy{138}{7}
\label{unpro1}
\end{eqnarray}
is unproductive: the special lifts of $239$, $378$ and $138$ are
respectively $8956$, $143640$ and $134683$, which happen to be in
arithmetic progression \mod{379^2}. The two weak 3AP decompositions
\begin{eqnarray}
\uu{11} &=& \cy{10}{2}\times\cy{1}{1}\times\cy{3}{5}
\label{unpro2}
\end{eqnarray}
and
\begin{eqnarray}
\uu{461} &=& \cy{1}{1}\times\cy{48}{4}\times\cy{95}{115}
\label{unpro3}
\end{eqnarray}
are also unproductive, but these are the only ones with prime modulus
less than $1000$. It appears that productive decompositions predominate;
unproductive ones depend on an accidental coincidence which is comparatively
rare.
\subparagraph{Subcase 2.2:} One of $a,b,c$ (say $a$, without loss) is
divisible by $p$. Now choose the special lift of either $y$ or $z$, and any
non-special lift of the other; the lift of $x$ is determined by the AP
requirement. So there are $2(p-1)$ lifts to 3AP decompositions.
Here is an example. Start from a weak 3AP decompositions of $\uu{55}$:
\begin{eqnarray}
\uu{55}=\cy{54}{2}\times\cy{1}{1}\times\cy{3}{20}\,.
\label{sct33}
\end{eqnarray}
We wish to lift to strong 3AP decompositions of $\uu{275}$. We are
in this subcase. All lifts of $3$ have order~$20$, but each of the
generators $54$ and $1$ has one special lift (namely $274$ and $1$
respectively). So we must use a non-special lift of $1$, the special
lift of $54$, and the lift of $3$ which completes the AP:
\begin{eqnarray*}
\uu{275} & = & \cy{274}{2} \times \cy{166}{5} \times \cy{58}{20} \\
& = & \cy{274}{2} \times \cy{56}{5} \times \cy{113}{20} \\
& = & \cy{274}{2} \times \cy{221}{5} \times \cy{168}{20} \\
& = & \cy{274}{2} \times \cy{111}{5} \times \cy{223}{20}
\end{eqnarray*}
The other four lifts (where we use the special lift of $1$ and a non-special
lift of~$54$) are weak 3AP decompositions. In the same way, the decomposition
\begin{eqnarray}
\uu{55} = \cy{52}{20} \times \cy{54}{2} \times \cy{1}{1}
\label{sct34}
\end{eqnarray}
gives rise to four more strong 3AP decompositions of $\uu{275}$.
Suppose, however, that we consider lifting (\ref{sct33}) and (\ref{sct34})
from $n = 55$ to \mbox{$n = 605$}. The only lifts of 52, 54, 1 and 3
\mod{605} which have orders 20, 2, 1, and 20 respectively are
602, 604, 1 and 3 \mod{605}, which are in AP, so we fail to obtain
any 3AP decomposition for $\uu{605}$.
\subparagraph{Subcase 2.3:} Two of $a,b,c$ (say $a$ and $b$) are divisible
by $p$. We must choose a non-special lift of $z$, and any lift of $x$; so
there are $p(p-1)$ lifts to 3AP decompositions of~$\uu{n'}$. Suppose,
for example, that we take $n = 273 = 3\times7\times13$ and $p=3$, to give
$n' = 819$. Computer enumeration has shown that there are $108$ strong
3AP decompositions of~$\uu{273}$, each perforce having 2 generators whose
orders are multiples of~3. The 648 lifts to $\uu{819}$ arise from the
strong decompositions.
\subparagraph{Subcase 2.4:} All three of $a,b,c$ are divisible by $p$. In
this case, three distinct prime divisors of $n$ are congruent to $1$ \mod{p},
and so $np$ has at least four prime divisors, so no
3AP decomposition of $\uu{np}$ can exist. (Alternatively, note that all
lifts of $x,y,z$ have the same orders as the original elements, so the group
they generate has the same order as $\langle x,y,z\rangle$.)
\paragraph{Case 3:} $p$ does not divide~$n$. In this case,
$\uu{np}\cong\uu{n}\times\uu{p}$. This case is the most difficult and
we do not have any general criteria for a lift to exist. However,
$\uu{np}$ is a product of at most 3 cyclic groups. From the structure
of the group of units, as described in the Introduction, we see that
$n=q^\alpha r^\beta$, or $2q^\alpha r^\beta$,
or $4q^\alpha$, or $2^\alpha$, for some odd primes $q$ and $r$, and some
$\alpha,\beta\ge0$. In this case, one of the lifts of any generator of a
3AP decomposition of $\uu{n}$ is a multiple of $p$, and therefore must be
disallowed as a \emph{spurious lift}. The spurious lifts $a',b',c'$ of the
three generators $a,b,c$ are in~AP. For they form an AP (mod~$n$) by
definition, and a trivial AP (mod~$p$). By the Chinese Remainder Theorem,
the congruences
\[b'-a'\equiv c'-b'\hbox{ (mod~$n$)},\qquad b'-a'\equiv c'-b'\hbox{ (mod~$p$)}\]
imply that $b'-a'\equiv c'-b'$ (mod~$pn$).
Suppose that we have $n = 31$ and $p = 5$, and we consider lifting
\[\uu{31} = \cy{25}{3}\times\cy{30}{2}\times\cy{4}{5}\,.\]
The respective spurious lifts are 25, 30 and 35. Two lifts of the
3AP decomposition are available:
\begin{eqnarray*}
\uu{155} & = & \cy{56}{3}\times\cy{154}{2}\times\cy{97}{20}\\
& & \cy{87}{12}\times\cy{154}{2}\times\cy{66}{5}\,.
\end{eqnarray*}
\subsection{Lifting to $\uu{p^\alpha}$}
\label{ss33}
In one special case of lifts we can draw a strong conclusion.
This case requires a productive 3AP decomposition for $\uu{p}$ where
$p$ is prime; we recall the unproductive examples (\ref{unpro1}),
(\ref{unpro2}) and (\ref{unpro3}) given above.
\begin{theorem}
Let $p$ be an odd prime, and suppose that
\[\uu{p}=\cy{x}{a}\times\cy{y}{b}\times\cy{z}{c}\]
is a productive (possibly weak) \textup{3AP} decomposition. Then for any
$\alpha\ge2$, there is a lift of the given decomposition which is a (strong)
\textup{3AP} decomposition of~$\uu{p^\alpha}$.
\end{theorem}
\begin{pf}
We show inductively that there is a lift where two of the lifted elements
have the same orders as the originals, and the third has order multiplied
by~$p^{\alpha-1}$.
For the first step, we are in case 2, subcase 2.1; in this case we saw
that any productive decomposition has three lifts, at least one of which
is strong.
For the general step, we are in case 1, and we start with a decomposition
in which two of the elements have orders coprime to $p$; so the necessary
condition for this case is satisfied, and the lift exists.
\end{pf}
\section{$n$ an odd composite integer}
\label{s4}
\subsection{$n$ a multiple of~$3$}
\label{ss41}
\begin{theorem}
A \textup{3AP} decomposition of $\uu{3p}$ does not exist for any prime $p$
with $p > 3$.
\label{t3p}
\end{theorem}
\begin{pf} Suppose that
\[\uu{3p}=\langle a\rangle\times\langle a+d\rangle\times\langle a+2d\rangle.\]
Then $d$ is divisible by $3$, since otherwise one of $a$, $a+d$, $a+2d$ would
be a multiple of~$3$.
If $a\equiv1$~(mod~$3$) then all three generators are congruent
to $1$~(mod~$3$), and so is every element in the group they generate, which
is not possible. On the other hand, if $a\equiv2$~(mod~$3$), then each of
the generators has even order (since it has even order in $\uu{3}$), and so
$C_2\times C_2\times C_2\le\uu{3p}=C_2\times C_{p-1}$, a contradiction.
\end{pf}
The next result has a similar proof; it is rather special but rules out one
particular type of 3AP decomposition.
\begin{theorem}
There is no \textup{3AP} decomposition of $\uu{3m}$ of the form
\[\uu{3m}=\langle a\rangle\times\langle a+m\rangle\times\langle a+2m\rangle.\]
\end{theorem}
\begin{pf}
The argument of the preceding theorem shows that $m$ is divisible by~$3$.
Since all the generators are congruent mod~$m$,
and projection from $\uu{3m}$ to $\uu{m}$ is onto, we see that $\uu{m}$ must
be cyclic (and $a$ is a primitive root of $m$), so $m$ is of the form $p^t$,
or $2p^t$ (for some odd prime $p$), or $m=4$. So necessarily $p=3$. But
now the order of $a$ (mod~$m$) is $2\cdot3^{t-1}$, and the same goes for the
other generators as well. Their orders (mod~$3m$) are at least as large, so
we must have $(2\cdot3^{t-1})^3\le 2\cdot3^t$, which is impossible.
\end{pf}
\subsection{Products of three primes}
\label{ss42}
Theorem~\ref{t3p} does not rule out 3AP decompositions of $\uu{3pq}$, where
$p$ and $q$ are distinct primes, and indeed these do exist. In this case,
a new phenomenon occurs: we can obtain new solutions from old. This works
more generally for the case where $n$ is the product of three odd primes
$p,q,r$, and $\xi(n)=4$ (so that $\uu{n}=C_{\lambda_n}\times C_2\times C_2$).
Suppose that the abelian group $A$ can be written (adapting our previous
notation) as
\[A=\cy{x}{2a}\times\cy{y}{2}\times\cy{z}{2}.\]
Then $A$ contains an elementary abelian group $B$ of of order~$8$ generated by
$x^a,y,z$. If three elements $x',y',z'$ have the properties that their orders
are $2a,2,2$ respectively and $\langle(x')^a,y',z'\rangle=B$, then $x',y',z'$
generate cyclic subgroups whose direct product is $A$.
If $A=\uu{n}$ for some $n$, then multiplying an
arithmetic progression by a fixed unit yields an arithmetic progression; so
we look for an element $u$ such that $x'=xu$, $y'=yu$, $z'=zu$ satisfy the
above conditions.
We see that $u$ must have order~$2$, so $u\in B$. If $a$ is even then
$(xu)^a=x^a$, while if $a$ is odd then $(xu)^a=x^au$. It is then easy to
check that the allowable values of $u$ are as follows:
\begin{itemize}
\item $u\in\{x^a,yz,x^ayz\}$ if $a$ is even;
\item $u\in\{x^ay,x^az,yz\}$ if $a$ is odd.
\end{itemize}
In each case, the possible values of $u$, together with the identity, form
a subgroup of $B$; so no further expressions can be obtained by repeating
the procedure. Moreover, in each case, $yz$ is an allowed multiplier, and
converts $[x,y,z]$ into $[xyz,z,y]$; so the solutions come in pairs, each
pair consisting of the first three and the last three terms in the sequence
$[x,y,z,xyz]$.
\begin{theorem}
Suppose that $n$ is the product of three odd primes, and that
\[\uu{n}=\cy{x}{\lambda}\times\cy{y}{2}\times\cy{z}{2}\]
is a \textup{3AP} decomposition, where $\lambda=\lambda_n$. Then
\[\uu{n}=\cy{ux}{\lambda}\times\cy{uy}{2}\times\cy{uz}{2}\]
is also a \textup{3AP} decomposition, where
\begin{itemize}
\item $u\in\{x^{\lambda/2},yz,x^{\lambda/2}yz\}$ if
$\lambda\equiv0\,(\mathrm{mod}\,4)$;
\item $u\in\{x^{\lambda/2}y,x^{\lambda/2}z,yz\}$ if
$\lambda\equiv2\,(\mathrm{mod}\,4)$.
\end{itemize}
In each case, there are two four-term arithmetic progressions whose
three-term subprogressions give the stated decompositions.
\end{theorem}
We call these sets of four decompositions \emph{quartets}.
Here are some examples of quartets, in cases where one of the primes
dividing $n$ is~$3$. We list the values of $n$ and $\lambda$, and the
two four-term progressions $[x,y,z,xyz]$; the orders of the terms are
$\lambda,2,2,\lambda$, and the first and last three give 3AP decompositions.
\paragraph{Case} $\lambda\equiv0$ (mod~$4$):
\begin{itemize}
\item $105$; $12$; $[38, 71, 104, 32]$, $[17, 29, 41, 53]$
\item $165$; $20$; $[113, 56, 164, 107]$, $[47, 89, 131, 8]$
\item $285$; $36$; $[98, 191, 284, 92]$, $[212, 134, 56, 263]$
\item $357$; $48$; $[122, 239, 356, 116]$, $[269, 50, 188, 326]$
\item $465$; $60$; $[158, 311, 464, 152]$, $[437, 404, 371, 338]$
\end{itemize}
\paragraph{Case} $\lambda\equiv2$ (mod~$4$):
\begin{itemize}
\item $231$; $30$; $[80, 155, 230, 74]$, $[179, 188, 197, 206]$
\item $483$; $66$; $[164, 323, 482, 158]$, $[95, 461, 344, 227]$
\end{itemize}
In general, there is no rquirement that an end-term in a quartet
should be the product \mod{n} of the other three terms. A
counter-example is the following, where the subscript integers
are the orders of the terms:
\begin{itemize}
\item $315$; $12$; $[8_4, 131_6, 254_6, 62_4]$
\end{itemize}
\subsection{Some results for $n = pq$ $(p > 3, q > 3)$}
We now indicate how the role of the value ${-3}$ \mod{n}, as discussed
in \S\ref{s2} above, carries over to composite values of $n$.
\begin{theorem}
Let $p$ and $q$ be primes greater than $3$, with $p\equiv3$ \mod{4}, and
suppose that $\ensuremath{\mathrm{ord}}_p(-3)=(p-1)/2$ and $\ensuremath{\mathrm{ord}}_q(-3)=q-1$.
(This implies that $p\equiv1$ \mod{3} and $q\equiv2$ \mod{3}).
Let $n=pq$, and let $x$ be the unique element of $\ensuremath{\mathbb{U}_n}$ congruent
to $1$ \mod{p} and to $-3$ \mod{q}. Then
\[\ensuremath{\mathbb{U}_n} = \langle-x-2\rangle_{(p-1)/2} \times \langle-1\rangle_2 \times
\langle x\rangle_{q-1}\,,\]
which is a lift of $\uu{p}=\cy{-3}{(p-1)/2}\times\cy{-1}{2}\times\cy{1}{1}$.
\label{t41}
\end{theorem}
\begin{pf}
The congruences \mod{3} arise by noticing that $-3$ is a quadratic residue
\mod{p} and non-residue \mod{q}, and applying quadratic reciprocity.
We have $\ensuremath{\mathrm{ord}}_p(x)=1$ and $\ensuremath{\mathrm{ord}}_q(x)=q-1$, so $\ensuremath{\mathrm{ord}}_n(x)=q-1$. Also, $-x-2$
is congruent to $-3$ \mod{p} and to $1$ \mod{q}, so $\ensuremath{\mathrm{ord}}_p(-x-2)=(p-1)/2$
and $\ensuremath{\mathrm{ord}}_q(-x-2)=1$, whence $\ensuremath{\mathrm{ord}}_n(-x-2)=(p-1)/2$.
Since $p\equiv3$ \mod{4}, we have $\uu{p} = \langle-3\rangle_{(p-1)/2}
\times \langle-1\rangle_2$, as the orders of the factors are co-prime.
So the group $A=\langle-x-2\rangle_{(p-1)/2} \times \langle-1\rangle_2 \times
\langle x\rangle_{q-1}$ projects onto $\mathbb{U}_p$. Also, $x$ belongs to
the kernel of this projection; since $x$ is a primitive root of $q$, the
kernel is $\mathbb{U}_q$. So $A=\ensuremath{\mathbb{U}_n}$.
\end{pf}
\paragraph{Coverage} In the range $n < 300$, the coverage of
Theorem~\ref{t41} is as follows:
\begin{eqnarray*}
35 = 7 \times 5: &&
\uu{35}=\cy{11}{3} \times \cy{34}{2} \times \cy{22}{4} \\
77 = 7 \times 11: &&
\uu{77}=\cy{67}{3} \times \cy{76}{2} \times \cy{8}{10} \\
95 = 19 \times 5: &&
\uu{95}=\cy{16}{9} \times \cy{94}{2} \times \cy{77}{4} \\
119 = 7 \times 17: &&
\uu{119}=\cy{18}{3} \times \cy{118}{2} \times \cy{99}{16} \\
155 = 31 \times 5: &&
\uu{155}=\cy{121}{15}\times \cy{154}{2} \times \cy{32}{4} \\
161 = 7 \times 23: &&
\uu{161}=\cy{116}{3} \times \cy{160}{2} \times \cy{43}{22} \\
203 = 7 \times 29: &&
\uu{203}=\cy{88}{3} \times \cy{202}{2} \times \cy{113}{28} \\
209 = 19 \times 11: &&
\uu{209}=\cy{111}{9} \times \cy{208}{2} \times \cy{96}{10} \\
215 = 43 \times 5: &&
\uu{215}=\cy{126}{21}\times \cy{214}{2} \times \cy{87}{4}
\end{eqnarray*}
The case $287=7\times41$ fails, since $\ensuremath{\mathrm{ord}}_{41}({-3}) = 8$.
In the range $q < 300$, the value $q = 41$ is the only prime $q$ with
$q \equiv 2$ \mod{3} and $\ensuremath{\mathrm{ord}}_q(-3) \neq q-1$. However, in the range
$p < 300$, there are four primes $p$ with $p \equiv 7$ \mod{12} and
$\ensuremath{\mathrm{ord}}_p(-3) \neq (p-1)/2$, namely $p = 67$, 103, 151 and 271.
\bigskip
As is hinted in \S8 of \cite{cap}, many special cases arise when
we come to consider composite values $n = pq$
where $p$ and $q$ are distinct primes satisfying
$p \equiv q \equiv 1$ \mod{6}, with \textup{gcd}$(p-1,\,q-1) = 6$.
Accordingly, we do not offer theorems to cover these cases.
Instead, for the range $n < 1000$, we use Table~1 to list
the instances in which we have
\begin{eqnarray}
\ensuremath{\mathbb{U}_n} = \cy{2x+3}{m} \times \cy{x+1}{3} \times \cy{-1}{2}
= \cy{-2x-3}{m} \times \cy{-x-2}{3} \times \cy{-1}{2}
\end{eqnarray}
where $m = \phi_n / 6$. The following values of $n$ are \textbf{not}
covered :
$259 = 7 \times 37$,\ \ $427 = 7 \times 61$,\ \
$511 = 7 \times 73$ and $973 = 7 \times 139$.
\begin{table}[p]
TABLE 1\\ \\
Decompositions (9) for $\ensuremath{\mathbb{U}_n}$ where $n = pq$ as specified in the text
\begin{center}
$
\begin{array}
{@{\hspace{1mm}}r@{\hspace{1mm}}c@{}c@{\hspace{2mm}}l@{\hspace{1mm}}l}
\hline
91&=& 7 \times 13: &
\uu{91} &=\cy{33}{12} \times \cy{16}{3} \times \cy{90}{2}
=\cy{58}{12} \times \cy{74}{3} \times \cy{90}{2} \\
133&=& 7 \times 19: &
\uu{133}&=\cy{61}{18} \times \cy{30}{3} \times \cy{132}{2}
=\cy{72}{18} \times \cy{102}{3} \times \cy{132}{2} \\
217&=& 7 \times 31: &
\uu{217}&=\cy{135}{30} \times \cy{67}{3} \times \cy{216}{2}
=\cy{82}{30} \times \cy{149}{3} \times \cy{216}{2} \\
247&=&13 \times 19: &
\uu{247}&=\cy{137}{36} \times \cy{68}{3} \times \cy{246}{2}
=\cy{110}{36} \times \cy{178}{3} \times \cy{246}{2} \\
& & &&=\cy{175}{36} \times \cy{87}{3} \times \cy{246}{2}
=\cy{72}{36} \times \cy{159}{3} \times \cy{246}{2} \\
301&=& 7 \times 43: &
\uu{301}&=\cy{271}{42} \times \cy{135}{3} \times \cy{300}{2}
=\cy{30}{42} \times \cy{165}{3} \times \cy{300}{2} \\
403&=&13 \times 31: &
\uu{403}&=\cy{228}{60} \times \cy{315}{3} \times \cy{402}{2}
=\cy{175}{60} \times \cy{87}{3} \times \cy{402}{2} \\
469&=& 7 \times 67: &
\uu{469}&=\cy{142}{66} \times \cy{305}{3} \times \cy{468}{2}
=\cy{327}{66} \times \cy{163}{3} \times \cy{468}{2} \\
553&=& 7 \times 79: &
\uu{553}&=\cy{205}{78} \times \cy{102}{3} \times \cy{552}{2}
=\cy{348}{78} \times \cy{450}{3} \times \cy{552}{2} \\
559&=&13 \times 43: &
\uu{559}&=\cy{202}{84} \times \cy{380}{3} \times \cy{558}{2}
=\cy{357}{84} \times \cy{178}{3} \times \cy{558}{2} \\
589&=&19 \times 31: &
\uu{589}&=\cy{547}{90} \times \cy{273}{3} \times \cy{588}{2}
=\cy{42}{90} \times \cy{315}{3} \times \cy{588}{2} \\
679&=& 7 \times 97: &
\uu{679}&=\cy{26}{96} \times \cy{352}{3} \times \cy{678}{2}
=\cy{653}{96} \times \cy{326}{3} \times \cy{678}{2} \\
721&=& 7 \times 103:&
\uu{721}&=\cy{422}{102} \times \cy{571}{3} \times \cy{720}{2}
=\cy{299}{102} \times \cy{149}{3} \times \cy{720}{2}\\
763&=& 7 \times 109:&
\uu{763}&=\cy{236}{108} \times \cy{499}{3} \times \cy{762}{2}
=\cy{527}{108} \times \cy{263}{3} \times \cy{762}{2}\\
& & &&=\cy{345}{108} \times \cy{172}{3} \times \cy{762}{2}
=\cy{418}{108} \times \cy{590}{3} \times \cy{762}{2}\\
817&=&19 \times 43:&
\uu{817}&=\cy{357}{126} \times \cy{178}{3} \times \cy{816}{2}
=\cy{460}{126} \times \cy{638}{3} \times \cy{816}{2}\\
871&=&13 \times 67:&
\uu{871}&=\cy{59}{132} \times \cy{29}{3} \times \cy{870}{2}
=\cy{812}{132} \times \cy{841}{3} \times \cy{870}{2}\\
& & &&=\cy{410}{132} \times \cy{640}{3} \times \cy{870}{2}
=\cy{461}{132} \times \cy{230}{3} \times \cy{870}{2}\\
889&=& 7 \times127:&
\uu{889}&=\cy{674}{126} \times \cy{781}{3} \times \cy{888}{2}
=\cy{215}{126} \times \cy{107}{3} \times \cy{888}{2}\\
\hline
\end{array}
$
\end{center}
\end{table}
Now let $n = pq$ where $p$ and $q$ are distinct primes satisfying
$p \equiv q \equiv 5$ \mod{8}, $q > 5$ and gcd$(p-1,q-1) = 4$.
For the range $n < 1000$, Table~2 lists 3AP decompositions of~$\ensuremath{\mathbb{U}_n}$
that are lifts from weak 3AP decompositions of~$\uu{q}$;
an asterisk marks a generator lifted from~1~\mod{q}.
Where the weak 3AP decomposition has a generator of order 4,
we classify the lifted 3AP decompos\-itions into three types:
if the generator lifted from~1 can be placed first, we
have type~A when the order of the middle generator is 4, and type~C
when the order of the last generator is 4, whereas type~B has the
generator lifted from~1 in the middle.
\begin{table}[p]
TABLE 2\\
Some lifts from weak 3AP decompositions of $\uu{q}$, as specified
at the end of \S4.3
\renewcommand{\arraystretch}{1.1}
\begin{center}
$
\begin{array}{rcc}
n = p \times q & \textup{3AP decomposition of } \ensuremath{\mathbb{U}_n} & \textup{Type} \\
\hline
65 = 5 \times 13 & \cy{27*}{4} \times \cy{44}{4} \times \cy{61}{3} & A \\
& \cy{53*}{4} \times \cy{57}{4} \times \cy{61}{3} & A \\
& \cy{53*}{4} \times \cy{16}{3} \times \cy{44}{4} & C \\
145 = 5 \times 29 & \cy{88*}{4} \times \cy{12}{4} \times \cy{81}{7} & A \\
& \cy{117*}{4} \times \cy{99}{4} \times \cy{81}{7} & A \\
185 = 5 \times 37 & \cy{43}{4} \times \cy{112*}{4} \times \cy{181}{9} & B \\
& \cy{38*}{4} \times \cy{16}{9} \times \cy{179}{4} & C \\
265 = 5 \times 53 & \cy{213*}{4} \times \cy{201}{13} \times \cy{189}{4} & C \\
305 = 5 \times 61 & \cy{123*}{4} \times \cy{56}{15} \times \cy{294}{4} & C \\
& \cy{62*}{4} \times \cy{24}{20} \times \cy{291}{3} & - \\
& \cy{273}{12} \times \cy{62*}{4} \times \cy{156}{5} & - \\
377 = 13\times 29 & \cy{262*}{12} \times \cy{99}{4} \times \cy{313}{7} & A \\
377 = 29\times 13 & \cy{287*}{28} \times \cy{57}{4} \times \cy{203}{3} & A \\
& \cy{14*}{28} \times \cy{146}{3} \times \cy{278}{4} & C \\
& \cy{222*}{28} \times \cy{146}{3} \times \cy{70}{4} & C \\
& \cy{235*}{28} \times \cy{146}{3} \times \cy{57}{4} & C \\
505 = 5 \times 101& \cy{102*}{4} \times \cy{394}{4} \times \cy{181}{25} & A \\
& \cy{102*}{4} \times \cy{414}{4} \times \cy{221}{25} & A \\
& \cy{203*}{4} \times \cy{192}{4} \times \cy{181}{25} & A \\
& \cy{203*}{4} \times \cy{212}{4} \times \cy{221}{25} & A \\
& \cy{203*}{4} \times \cy{56}{25} \times \cy{414}{4} & C \\
545 = 5 \times 109& \cy{33}{4} \times \cy{437*}{4} \times \cy{296}{27} & B \\
& \cy{403}{4} \times \cy{437*}{4} \times \cy{471}{27} & B \\
689 = 13\times 53 & \cy{319*}{12} \times \cy{625}{13} \times \cy{242}{4}& C \\
689 = 53\times 13 & \cy{209*}{52} \times \cy{317}{4} \times \cy{425}{3} & A \\
& \cy{469*}{52} \times \cy{447}{4} \times \cy{425}{3} & A \\
& \cy{456*}{52} \times \cy{107}{3} \times \cy{447}{4} & C \\
& \cy{586*}{52} \times \cy{107}{3} \times \cy{317}{4} & C \\
745 = 5 \times 149& \cy{193}{4} \times \cy{597*}{4} \times \cy{256}{37} & B \\
& \cy{403}{4} \times \cy{597*}{4} \times \cy{46}{37} & B \\
785 = 5 \times 157& \cy{158*}{4} \times \cy{757}{4} \times \cy{571}{39} & A \\
& \cy{472*}{4} \times \cy{129}{4} \times \cy{571}{39} & A \\
& \cy{443}{4} \times \cy{472*}{4} \times \cy{501}{39} & B \\
& \cy{158*}{4} \times \cy{207}{12} \times \cy{256}{13}& - \\
& \cy{158*}{4} \times \cy{326}{3} \times \cy{494}{52} & - \\
865 = 5 \times 173& \cy{693*}{4} \times \cy{566}{43} \times \cy{439}{4} & - \\
905 = 5 \times 181& \cy{363*}{4} \times \cy{316}{5} \times \cy{269}{36} & - \\
985 = 5 \times 197& \cy{183}{4} \times \cy{592*}{4} \times \cy{16}{49} & B \\
\hline
\end{array}
$
\end{center}
\renewcommand{\arraystretch}{1.4}
\end{table}
\subsection{Lifts from $n = kp$ to $n = kp^2$}
For the range $n < 1000$, details of the 3AP decompositions (3APDs) for
values of the form $n = kp^2$ ($k$ and $p$ distinct odd primes, $k > 3$,
$p > 3$) are as in Table~3.
With the given restrictions on $k$ and $p$,
just one value of the form $n = kp^3$ lies in the range $n < 1000$,
namely $n = 875$, and it has precisely six 3AP decompositions.
Each of these is obtained by further lifting one of the 3AP decompositions
for $n = 175$. In this further lifting, the orders that are not
multiples of 5 are unchanged, but the orders that are multiples of 5
become multiples of $5^2$.
\begin{table}[p]
TABLE 3\\ \\
An enumeration of decompositions for $n = kp^2$\\
\begin{center}
\renewcommand{\arraystretch}{1.1}
$
\begin{array}{lr@{\hspace{5mm}}rrr}
&& \multicolumn{3}{c}{\overbrace{\hspace{60mm}}} \\
\ \ \ n = kp^2 & \#\textup{ 3APDs}
& \multicolumn{2}{c}{\#\textup{ lifts from }\uu{kp}}
& \#\textup{ other} \\
\cline{3-4}
&& \multicolumn{1}{c}{\textup{from}}
& \multicolumn{1}{c}{\textup{from}} & \\
&& \textup{strong 3APDs}
& \textup{weak 3APDs} & \\
\hline
175 = 7 \cdot 5^2 & 6 & 3 & 3 & 0 \\
245 = 5 \cdot 7^2 & 6 & 3 & 3 & 0 \\
275 = 11 \cdot 5^2 & 68 & 0 & 8 & 60 \\
325 = 13 \cdot 5^2 & 20 & 12 & 8 & 0 \\
425 = 17 \cdot 5^2 & 8 & 0 & 8 & 0 \\
475 = 19 \cdot 5^2 & 6 & 3 & 3 & 0 \\
539 = 11 \cdot 7^2 & 12 & 9 & 3 & 0 \\
575 = 23 \cdot 5^2 & 2 & 0 & 2 & 0 \\
605 = 5 \cdot 11^2 & 0 & 0 & 0\rlap{*} & 0 \\
637 = 13 \cdot 7^2 & 126 & 108 & 18 & 0 \\
725 = 29 \cdot 5^2 & 30 & 18 & 12 & 0 \\
775 = 31 \cdot 5^2 & 188 & 32 & 24 & 132 \\
845 = 5 \cdot 13^2 & 20 & 12 & 8 & 0 \\
847 = 7 \cdot 11^2 & 0 & 0\rlap{*} & 0\rlap{*} & 0 \\
925 = 37 \cdot 5^2 & 10 & 6 & 4 & 0 \\
931 = 19 \cdot 7^2 & 182 & 156 & 26 & 0 \\
\hline\\
\end{array}
$\\
\renewcommand{\arraystretch}{1.4}
{*} 3APDs of $\uu{kp}$ exist, but the special lifts of the generators
are in AP
\end{center}
\end{table}
As Table~3 indicates, some of the 3AP decompositions for $n = 275$
and $775$ are not lifts, these two $n$-values being distinctive in that
they have $ p\ |\ (k-1) $\,. How do these exceptional 3AP decompositions
arise? One of them is given by
\[ \uu{275} = \cy{16}{5} \times \cy{24}{10} \times \cy{32}{4}\,. \]
This is related to the decomposition
\[ \uu{55} = \phantom{\cy{116}{5} \times} \cy{24}{10} \times \cy{32}{4} \]
and to the fact that, within $\uu{55}$, we have
$\cy{16}{5} \subset \cy{24}{10}$. The further 3AP decomposition
\[ \uu{275} = \cy{181}{5} \times \cy{244}{10} \times \cy{32}{4} \]
arises in the same way, as 181 is a lift of 16, and 244 is a lift of 24.
Less straightforward situtations exist too. Consider, for example, the
3AP decomposition
\[ \uu{775} = \cy{32}{4} \times \cy{54}{10} \times \cy{76}{15} \]
with $n = 31 \cdot 5^2$. If we try lifting to this from
$n = 31 \cdot 5 = 155$,
we find that 32, 54 and 76 also have orders 4, 10 and 15 \mod{155}, and that
\begin{eqnarray*}
\uu{155} & = & \cy{32}{4} \times \cy{54}{10} \times \cy{76^5}{3} \\
& = & \cy{32}{4} \times \cy{54^5}{2} \times \cy{76}{15}\,.
\end{eqnarray*}
Analogous to this, we can rewrite an example from the previous paragraph
in the weak form
\[ \uu{55} = \cy{16^5}{1} \times \cy{24}{10} \times \cy{32}{4}\,. \]
These examples suggest an amusing generalisation of 3AP decompositions to
decompositions of the form
\[ \ensuremath{\mathbb{U}_n} = \cy{x^h}{} \times \cy{(x+k)^i}{} \times \cy{(x+2k)^j}{} \]
where $h,\,i,\,j \ge 1\,$, but we do not pursue this idea further here.
\section{A class of weak 3AP decompositions}
\label{s5}
A noteworthy class of weak 3AP decompositions arises for primes $n$
satisfying $n \equiv 1$ \mod{6p} where $p$ is an odd prime, $p > 3$.
In each of these decompositions, one of the generators has order 6 and
another has order~$p$. For $n < 300$, such decompositions are as follows:
\begin{eqnarray*}
n = 43 && \cy{1}{1}\times\cy{4}{7}\times\cy{7}{6} \\
n = 67 && \cy{1}{1}\times\cy{30}{6}\times\cy{59}{11} \\
n = 79 && \cy{1}{1}\times\cy{52}{13}\times\cy{24}{6} \\
n = 103 && \cy{1}{1}\times\cy{47}{6}\times\cy{93}{17} \\
n = 139 && \cy{97}{6}\times\cy{1}{1}\times\cy{44}{23} \\
n = 223 & \left\{ \begin{array}{@{}c@{}} \\ \\ \end{array} \right.
& \begin{array}{@{}l@{}}
\cy{1}{1}\times\cy{132}{37}\times\cy{40}{6} \\
\cy{184}{6}\times\cy{1}{1}\times\cy{41}{37} \\
\end{array} \\
n = 283 && \cy{45}{6}\times\cy{1}{1}\times\cy{240}{47}
\end{eqnarray*}
Where $\ensuremath{\mathbb{U}_n} = \cy{a}{6}\times\cy{1}{1}\times\cy{c}{p}$, we have
$a \equiv (c-1)^{-1}$ \mod{n}, so that
$\ensuremath{\mathbb{U}_n} = $ \mbox{$\cy{c-1}{6}\times\cy{c}{p}$}, a situation discussed in
\cite[\S8.2]{cap}. The above weak 3AP decomposition for $n = 67$
has the \emph{orders} of the generators in AP.
\section{Finite fields}
\label{s6}
Finite fields of non-prime order can have 3AP decompositions.
Clearly this is impossible in fields of characteristic~$2$,
which contain no 3-term arithmetic progressions.
\paragraph{Example}
Using GAP~\cite{GAP}, we found the following 3AP decompositions
of small finite fields $\mathop{\mathrm{GF}}(q)$.
In this list, $\zeta$ denotes the primitive root
(denoted by \verb+Z(q)+ in GAP notation) in the field $\mathop{\mathrm{GF}}(q)$.
It is a root of the appropriate \emph{Conway polynomial} \cite{Conway};
the relevant Conway polynomials are as follows:
\begin{eqnarray*}
q=11^2: && x^2+7x+2 \\
q=11^3: && x^3+2x+9 \\
q=19^2: && x^2-x+2 \\
q=19^3: && x^3+4x-2 \\
q=23^2: && x^2-2x+5 \\
q=29^2: && x^2-5x+2
\end{eqnarray*}
We give only
one decomposition for each possible list of orders of the factors:
\begin{eqnarray*}
\mathop{\mathrm{GF}}(11^2)^\times
&=& \cy{\zeta^{72}}{5} \times \cy{\zeta^{15}}{8} \times \cy{\zeta^{80}}{3} \\
\mathop{\mathrm{GF}}(11^3)^\times
&=& \cy{\zeta^{570}}{7} \times \cy{\zeta^{532}}{5} \times \cy{\zeta^{595}}{38} \\
&=& \cy{\zeta^{665}}{2} \times \cy{\zeta^{1008}}{95} \times \cy{\zeta^{570}}{7} \\
\mathop{\mathrm{GF}}(19^2)^\times
&=& \cy{\zeta^{144}}{5} \times \cy{\zeta^{320}}{9} \times \cy{\zeta^{135}}{8} \\
\mathop{\mathrm{GF}}(19^3)^\times
&=& \cy{\zeta^{3429}}{2} \times \cy{\zeta^{2970}}{127} \times \cy{\zeta^{5588}}{27} \\
\mathop{\mathrm{GF}}(23^2)^\times
&=& \cy{\zeta^{176}}{3} \times \cy{\zeta^{192}}{11} \times \cy{\zeta^{429}}{16} \\
\mathop{\mathrm{GF}}(29^2)^\times
&=& \cy{\zeta^{280}}{3} \times \cy{\zeta^{720}}{7} \times \cy{\zeta^{609}}{40}\\
&=& \cy{\zeta^{120}}{7} \times \cy{\zeta^{504}}{5} \times \cy{\zeta^{385}}{24}
\end{eqnarray*}
Can we have a 3AP decomposition of
$\mathop{\mathrm{GF}}(q)^\times$ in which two of the generators have orders $2$ and $3$\,?
As earlier, such a decomposition requires that $(q-1)/6$ is co-prime to $6$,
so that $q\equiv7$ or $31$ \mod{36}. But this implies that, if $q=p^n$ with
$p$ prime, then $p\equiv7$ or $31$ \mod{36} (since $7$ and $31$ are
primitive $\lambda$-roots\ of 36 \cite{cap}, and each is the fifth power of the other). Then
elements of orders $2$ and $3$ lie in the prime subfield, and hence
so does the whole AP. So there are no such decompositions
other than those of $\ensuremath{\mathbb{U}_n}$ for $n$ prime discussed in \S\ref{s2}.
A similar argument shows that a 3AP decomposition of $\mathop{\mathrm{GF}}(11^3)^\times$
into \mbox{factors} of orders $2$, $5$ and $133$ is impossible.
\section{Decompositions with more than three factors}
\label{s7}
As indicated above, we can define 4AP decompositions analogously
to 3AP decompositions.
A computer program has shown that
no examples of strong 4AP decompositions of $\ensuremath{\mathbb{U}_n}$ exist
for prime values of $n$ up to $10000$. The smallest composite $n$
for which strong 4AP decompositions of $\ensuremath{\mathbb{U}_n}$ exist is even:
\begin{eqnarray*}
\uu{104} & = &
\cy{31}{4}\times\cy{81}{3}\times\cy{27}{2}\times\cy{77}{2} \\
& = &
\cy{77}{2}\times\cy{79}{2}\times\cy{81}{3}\times\cy{83}{4}\,.
\end{eqnarray*}
The smallest weak $4$AP decomposition of $\ensuremath{\mathbb{U}_n}$ with prime $n$ is
\[\uu{3613} =
\cy{3528}{4} \times \cy{1148}{129} \times \cy{2381}{7} \times\cy{1}{1}\,.\]
We have no examples with larger numbers of generators that are in AP.
\paragraph{Note} The computations reported in this paper were performed
using GAP~\cite{GAP}, and a package of GAP functions written by the first
author for computations in the groups \ensuremath{\mathbb{U}_n}, available from~\cite{plrfns}.
Further documentation of these functions can be found in~\cite{cap}.
|
\section{Introduction}
There is hope that hard exclusive scattering processes in Bjorken kinematics
can provide one with a three-dimensional picture of the proton in
longitudinal and transverse plane~\cite{Burkardt:2000za}, encoded in generalized parton distributions
(GPDs)~\cite{Diehl:2003ny,Belitsky:2005qn}. One of the most important reactions in this context
is Compton scattering with one real and one highly-virtual photon (DVCS) which has
received a lot of attention. The QCD description of DVCS is based on the operator product
expansion (OPE) of the time-ordered product of two electromagnetic currents.
In this language the GPDs appear as leading-twist operator matrix elements.
In order to probe the transverse proton structure one needs to measure
the dependence of the amplitude on the momentum transfer to the target
$t=(P'-P)^2$ in a broad range. Since the available photon
virtualities $Q^2$ are limited to a few GeV$^2$ range,
corrections of the type $\propto t/Q^2$ (which are formally higher-twist effects),
are significant and have to be taken into account.
Such corrections are usually dubbed ``kinematic'' since they
only involve ratios of kinematic variables and at first sight have nothing to do
with nonperturbative effects (e.g. one may consider a theoretical
limit $\Lambda_{\rm QCD}^2 \ll t \ll Q^2$). The separation of
kinematic corrections $\propto t/Q^2$ from generic twist-four corrections
$\mathcal{O}(\Lambda_{\rm QCD}^2/Q^2)$ proves, however, to be surprisingly difficult.
The problem is well known and its importance for phenomenology has been
acknowledged by many
authors~\cite{Belitsky:2005qn,Blumlein:2000cx,Radyushkin:2000ap,Belitsky:2000vx,Kivel:2000rb,Belitsky:2001hz,Belitsky:2010jw,Geyer:2004bx,Blumlein:2006ia,Blumlein:2008di}.
The challenge is that, unlike target mass corrections in
inclusive reactions \cite{Nachtmann:1973mr}, which are determined
solely by the contributions of leading twist operators, the $\sim t/Q^2$ corrections
to off-forward processes (and for spin-1/2 targets also $\sim m^2/Q^2$ corrections)
also receive contributions from higher-twist-four operators that can be reduced
to total derivatives of the twist-two ones. Indeed,
let $\mathcal{O}_{\mu_1\ldots\mu_n}$ be a multiplicatively renormalizable
(conformal) local twist-two operator,
symmetrized and traceless over all indices. The operators
\begin{equation}
\mathcal{O}_1 = \partial^2 \mathcal{O}_{\mu_1\ldots\mu_n}\,, \qquad
\mathcal{O}_2 = \partial^{\mu_1}\mathcal{O}_{\mu_1\ldots\mu_n}
\label{eq:O1O2}
\end{equation}
are, on the one hand, twist-four, and on the other hand their matrix elements
are related to the leading twist matrix elements times the
momentum transfer squared (up to, possibly, target mass corrections).
Thus, both operators contribute to the $\propto t/Q^2$, $\propto m^2/Q^2$ accuracy
and must be taken into account.
Moreover, all these contributions are intertwined by electromagnetic gauge and
Lorentz invariance.
Implementation of the electromagnetic gauge invariance beyond the
leading twist accuracy has been at the center of
many discussions, starting from Ref.~\cite{Anikin:2000em}.
By contrast, importance of the translation invariance condition
has never been emphasized, to the best of our knowledge.
In particular the distinction
between the kinematic corrections of Nachtmann's type, i.e.
due to contributions of
leading-twist~\cite{Radyushkin:2000ap,Belitsky:2001hz,Belitsky:2000vx,Belitsky:2010jw
Geyer:2004bx,Blumlein:2006ia,Blumlein:2008di}, and of higher-twist operators in
Eq.~(\ref{eq:O1O2}) is not invariant under translations along the line connecting the
electromagnetic currents in the $T$-product. Hence
this distinction has no physical meaning; the existing estimates of kinematic effects,
e.g. in DVCS, by the contributions of leading twist operators alone can be misleading.
On a more technical level, the problem arises because $\mathcal{O}_2$ has rather
peculiar properties: the divergence of a conformal operator vanishes in the free theory
(the Ferrara-Grillo-Parisi-Gatto theorem \cite{Ferrara:1972xq}).
A related feature is that using QCD equations of motion (EOM) $\mathcal{O}_2$ can be
expressed in terms of quark-antiquark-gluon
operators. The simplest example of such a relation
is known for many years~\cite{Kolesnichenko:1984dj,Braun:2004vf,Anikin:2004ja}:
\begin{equation}
\partial^\mu O_{\mu\nu} = 2\bar q ig G_{\nu\mu}\gamma^\mu q\,,
\label{eq:puzzle}
\end{equation}
where
$
O_{\mu\nu} = (1/2)[\bar q \gamma_\mu\!\stackrel{\leftrightarrow}{D}_\nu \!q
+ (\mu\leftrightarrow\nu)]
$
is the quark part of the energy-momentum tensor. The operator on the r.h.s.
of Eq.~(\ref{eq:puzzle}) involves the gluon field strength and, naively,
its hadronic matrix elements are of the order of $\Lambda_{\rm QCD}^2$, which is in fact not the case.
More complicated examples can be found in \cite{Balitsky:1989ry,Ball:1998ff}.
The general structure of such relations is, schematically
\begin{equation}
(\partial\mathcal{O})_{N} = \sum_k a^{(N)}_{k}\,{G}_{Nk}\,,
\label{eq:partialO}
\end{equation}
where ${G}_{Nk}$ are twist-four quark-antiquark-gluon operators
and $a^{(N)}_{k}$ are the numerical coefficients.
The subscript $N$ stands for the number of
derivatives in $\mathcal{O}_N$ and the summation
goes over all contributing operators which may include total derivatives
(so that in practice $k$ is a certain multi-index).
The same operators, ${G}_{Nk}$, also appear in the OPE for
the product of currents of interest at the twist-four level:
\begin{equation}
T\{j(x)j(0)\}^{t=4} = \sum_{N,k} c_{N,k}(x)\,{G}_{Nk}\,.
\label{eq:Tproduct}
\end{equation}
A separation of ``kinematic'' and ``dynamical'' contributions to the OPE implies
that one attempts to reassemble this expansion in such a way that the contribution
of a particular combination appearing in (\ref{eq:partialO}) is
separated from the remaining twist-four contributions.
The ``kinematic'' power correction would correspond to taking into account
this term only, and discarding contributions of ``genuine'' quark-gluon operators.
The guiding principle is that the separation of kinematic and dynamical
effects is only physically meaningful (e.g. they are separately gauge- and Lorentz-invariant)
if they have autonomous scale dependence.
Different twist-four operators of the same dimension mix with each other
and satisfy a certain renormalization group (RG) equation which can be solved,
at least in principle. Let $\mathcal{G}_{N,k}$ be the set of multiplicatively
renormalizable twist-four operators so that
\begin{equation}
\mathcal{G}_{N,k} = \sum_{k'} \psi^{(N)}_{k,k'}\, {G}_{N,k'}\,.
\end{equation}
Eq.~(\ref{eq:partialO}) tells us that one of the solutions
of the RG equation is known {\it without calculation}. Indeed, it provides one with
an explicit expression for a twist-four operator with the anomalous dimension
equal to the anomalous dimension of the leading twist operator.
(For simplicity we ignore the contributions of $\partial^2\mathcal{O}_N$
in this discussion; they do not pose a problem and can
be taken into account using conventional methods.)
Let us assume that this special solution corresponds to $k=0$, i.e.
$\mathcal{G}_{N,k=0} \equiv (\partial\mathcal{O})_{N}$ and
$\psi^{(N)}_{k=0,k'}= a_{k'}$. Inverting the
matrix of coefficients, $\psi^{(N)}_{k,k'}$, and separating the term with $k=0$
we can write the expansion of an
arbitrary twist-four operator in terms of the multiplicatively renormalizable
ones
\begin{equation}
G_{N,k} = \phi^{(N)}_{k,0} (\partial\mathcal{O})_{N} + \sum_{k'\not=0}\phi^{(N)}_{k,k'}\,
\mathcal{G}_{N,k'}\,.
\label{eq:eee}
\end{equation}
Inserting this expansion into Eq.~(\ref{eq:Tproduct}) one obtains
\begin{equation}
T\{j(x)j(0)\}^{\rm tw-4} = \sum_{N,k} c_{N,k}(x)\phi^{(N)}_{k,0}\,(\partial\mathcal{O})_{N}
+ \ldots\,,
\label{eq:solve}
\end{equation}
where the ellipses stand for the ``genuine'' twist-four quark gluon
operators (e.g. with different anomalous dimensions).
This is the solution we want to have, but the problem with it is that
finding the coefficients $\phi^{(N)}_{k,0}$ in general requires knowledge of the full
matrix $\psi^{(N)}_{k,k'}$, in other words the explicit solution
of the twist-four RG equations, which is not available.
Our starting observation is that twist-four operators in QCD come
in two big groups: the so-called
quasipartonic~\cite{Bukhvostov:1985rn}, that only involve ``plus''
components of the fields, and non-quasipartonic which also include
``minus'' light-cone projections.
Quasipartonic operators are not relevant for the present discussion
since they have an autonomous evolution (to one-loop accuracy).
As a consequence, $(\partial\mathcal O)_N$ does not appear in the
expansion of quasipartonic operators in multiplicatively renormalizable ones, Eq.~(\ref{eq:eee}):
the corresponding coefficients
$\phi^{(N)}_{k,0}$ vanish. Hence the kinematic power correction
$\sim (\partial\mathcal O)_N$ originates entirely from contributions of
non-quasipartonic operators.
Renormalization of twist-four non-quasipartonic operators was
studied recently in~\cite{Braun:2008ia,Braun:2009vc}.
The main result is that in a suitable operator basis
the corresponding RG equations can be written in terms
of several $SL(2)$-invariant kernels.
Using $SL(2)$-invariance we are able to prove that
the anomalous dimension matrix for non-quasipartonic operators is
hermitian with respect to a certain scalar product. This implies that
different eigenvectors are mutually orthogonal, i.e.
\begin{equation}
\sum_k \mu^{(N)}_k \psi^{(N)}_{l,k} \psi^{(N)}_{m,k} \sim \delta_{l,m}\,,
\end{equation}
where $\mu^{(N)}_k$ is the corresponding (nontrivial) measure.
{}From this orthogonality relation and the expression (\ref{eq:partialO})
for the relevant eigenvector one obtains, for the non-quasipartonic
operators
\begin{equation}
\phi^{(N)}_{k,0} = a^{(N)}_{k} ||a^{(N)}||^{-2}\,,
\label{eq:coc}
\end{equation}
where $||a^{(N)}||^2 = \sum_k \mu^{(N)}_k (a^{(N)}_{k})^2$.
Inserting this expression into (\ref{eq:solve}) one ends up with
the desired separation of kinematic effects.
The actual derivation is done using the two-component
spinor formalism in intermediate steps and requires some specific techniques
of the $SL(2)$ representation theory.
This talk is based on the results presented in Ref.~\cite{Braun:2011zr};
details of the derivation will be given in a forthcoming paper.
\section{T-product of two electromagnetic currents}
We have been able to find the contributions related to the
leading-twist operator~(\ref{Olt}) in the $T$-product of two electromagnetic currents
$
T_{\mu\nu}=i\,T \{j_\mu^{em}(x) j_\nu^{em}(0)\}
$
to twist-four accuracy. The result can be brought to the form
\begin{align}\label{Tmn}
T_{\mu\nu} &= -\frac{1}{\pi^2x^4}\Big\{
x^{\alpha}\Big[S_{\mu\alpha\nu\beta} \mathbb{V}^\beta
+i\epsilon_{\mu\nu\alpha\beta} \mathbb{A}^\beta
\Big]
+x^2\Big[(x_\mu\partial_\nu +x_\nu\partial_\mu) \mathbb{X}
+(x_\mu\partial_\nu-x_\nu\partial_\mu) \mathbb{Y}
\Big]
\Big\}\,,
\end{align}
where $\partial_\mu = \partial/\partial x^\mu$,
$S_{\mu\alpha\nu\beta}=g_{\mu\alpha}g_{\nu\beta}+g_{\nu\alpha}g_{\mu\beta}-g_{\mu\nu}g_{\alpha\beta}$
and a totally antisymmetric tensor is defined such that $\epsilon_{0123}=1$. The
expansion of invariant functions $\mathbb{V}_\beta$ and $\mathbb{A}_\beta$ starts from twist two,
wheareas $\mathbb{X}$ and $\mathbb{Y}$ are already twist-four.
In order to write the result we first need to introduce some notations.
We define nonlocal (light-ray) vector $O_V$ and axial-vector $O_A$ operators of the leading-twist-two
as the generating functions for local twist-two operators
\begin{align}\label{Olt}
O(z_1x,z_2x) =& \big[\bar q(z_1 x)\slashed{x}\,(\gamma_5)\, Q^2\,q(z_2 x )\big]_{l.t.}.
\end{align}
Here $x_\mu$ is an arbitrary four-vector (not necessarily light-like), $z_1$ and $z_2$ are
real numbers and $Q$ is the matrix of quark electromagnetic charges.
Here and below the Wilson line between the quark fields is implied.
The leading-twist projector $[\ldots]_{l.t.}$ stands for the subtraction of traces
of the local operators so that by definition
\begin{eqnarray}
\lefteqn{
\big[\bar q(z_1 x)\slashed{x}\, Q^2\,q(z_2x)\big]_{l.t.} =}
\nonumber\\
&=& \sum_{N} \frac{1}{N!} x_\mu x_{\mu_1}\ldots x_{\mu_N} \Big\{\bar q(0)\gamma_\mu
[z_1\!\stackrel{\leftarrow}{D}_{\mu_1}+z_2\!\stackrel{\rightarrow}{D}_{\mu_1}]
\ldots
[z_1\!\stackrel{\leftarrow}{D}_{\mu_N}+z_2\!\stackrel{\rightarrow}{D}_{\mu_N}]Q^2 q(0)
-{\rm traces}\Big\}.
\end{eqnarray}
The leading-twist light-ray operators satisfy the Laplace equation
$
\partial_x^2 O(z_1x,z_2x) = 0\,.
$
The explicit form of the projector $[\ldots]_{l.t.}$ is irrelevant
for what follows. Useful representations can be found
e.g. in~\cite{Belitsky:2001hz,Balitsky:1987bk}.
Thanks to crossing symmetry the vector and axial-vector operators always appear to be antisymmetrized
and symmetrized over the quark and antiquark positions, respectively, so we
define the corresponding combinations:
\begin{eqnarray}
{O}^{(-)}_{V}(z_1,z_2)&=\!&
\big[\bar q(z_1 x)\slashed{x}\, Q^2\,q(z_2 x )\big]_{l.t.} \!- (z_1\leftrightarrow z_2)\,,
\\
{O}^{(+)}_{A}(z_1,z_2)&=\!&
\big[\bar q(z_1 x)\slashed{x}\,\gamma_5\, Q^2\,q(z_2 x )\big]_{l.t.} \!+ (z_1\leftrightarrow z_2)\,.
\nonumber
\end{eqnarray}
The leading-twist expressions are well known and can be written as
(cf.~\cite{Balitsky:1987bk})
\begin{equation}
\mathbb{V}^{t=2}_\mu
=\frac12 \partial_\mu
\int_0^1\!{du}\,{O}^{(-)}_{V}(u,0)\,,
\qquad\qquad
\mathbb{A}^{t=2}_\mu
=\frac12\partial_\mu
\int_0^1\!{du}\,{O}^{(+)}_{A}(u,0)\,.
\end{equation}
Note that the separation of the leading-twist terms $[\ldots]_{l.t.}$ from
the nonlocal operators produces a series of kinematic power corrections
to the amplitudes, which are similar to Nachtmann target mass corrections
in deep-inelastic lepton-nucleon scattering~\cite{Nachtmann:1973mr}. Such corrections
are discussed in detail
in~\cite{Kivel:2000rb,Belitsky:2001hz,Belitsky:2000vx,Belitsky:2010jw,Geyer:2004bx,Blumlein:2006ia,Blumlein:2008di}.
For the twist-three functions we obtain
\begin{align}\label{VV3}
\mathbb{V}^{t=3}_\mu=&
\Big[i\mathbf{P}^\nu,\int_0^1 \!\!du\,\Big\{i\epsilon_{\mu\alpha\beta\nu}
x^\alpha \partial^\beta\widetilde{{O}}^{(+)}_{A}(u)
+
\Big(S_{\mu\alpha\nu\beta}x^\alpha\partial^\beta+
\ln u \,\partial^\mu x^2\partial^\nu\Big)
\widetilde{{O}}^{(-)}_{V}(u)\Big\}\Big]\,,
\notag\\
\mathbb{A}^{t=3}_\mu=&
\Big[i\mathbf{P}^\nu,\int_0^1 \!\!du\,\Big\{i\epsilon_{\mu\alpha\beta\nu}
x^\alpha \partial^\beta\widetilde{{O}}^{(-)}_{V}(u)
+
\Big(S_{\mu\alpha\nu\beta}x^\alpha\partial^\beta+
\ln u \,\partial^\mu x^2\partial^\nu\Big)
\widetilde{{O}}^{(+)}_{A}(u)\Big\}\Big]\,.
\end{align}
Here $\mathbf{P}_\nu$ is the momentum operator
$
[i\mathbf{P}_{\!\!\nu}, q(y)]=\frac{\partial}{\partial y^\nu} q(y),
$
and we used the notation
\begin{equation}\label{Ot}
\widetilde{{O}}^{(\pm)}_a(z)=\frac{1}{4}\int_{0}^{z}\! dw \,{O}^{(\pm)}_a(z,w)\,.
\end{equation}
One can easily verify that $x^\mu \mathbb{V}_\mu^{t=3}=\partial^\mu \mathbb{V}_\mu^{t=3}=0$
and similarly $x^\mu \mathbb{A}_\mu^{t=3}=\partial^\mu \mathbb{A}_\mu^{t=3}=0$.
Note that the terms in $\ln u $ in Eqs.~(\ref{VV3}) are themselves twist-four
and can be omitted if the calculation is done to twist-three accuracy.
The resulting simplified expression is in
agreement with Refs.~\cite{Radyushkin:2000ap,Belitsky:2000vx}. These terms must be
included, however, in order to ensure correct separation of twist-three and twist-four
contributions.
The flavor-nonsinglet twist-four contributions to Eq.~(\ref{Tmn})
present our main result. In this case we prefer to write the answer in
terms of integrals over the position of the local conformal operators, cf. Eq.~(\ref{eq:COPE}).
This form is usually referred to as the conformal OPE~\cite{Braun:2003rp}.
For example, a light-ray operator can be written as the conformal expansion
\begin{equation}
{O}(z_1x,z_2x) =\!\sum_N\varkappa_N\, z_{12}^N\int_0^1\! du\, (u\bar u)^{N+1}
\bigl[ \mathcal{O}_{N}(z_{12}^ux)\bigr]_{l.t.}\,,
\label{eq:COPE}
\end{equation}
where $$\varkappa_N=2(2N+3)/(N+1)!$$ and we use the shorthand notation
$ \bar u =1-u\,,\quad z_{12}=z_1-z_2\,,\quad z^u_{12} = \bar u z_1 + u z_2$.
The conformal operator $\mathcal{O}_{N}$ is defined as
\begin{align}
\mathcal{O}_{N}(y)=&
(\partial_{z_1}\!+\!\partial_{z_2})^NC_N^{3/2}\left(
\frac{\partial_{z_1}\!-\!\partial_{z_2}}{\partial_{z_1}\!+\!\partial_{z_2}}\right)
\,{O}(z_1x +y,z_2x+y)\Big|_{z_i=0},
\label{eq:On}
\end{align}
where $C_N^{3/2}(x)$ is the Gegenbauer polynomial.
The leading-twist contribution to the OPE of two electromagnetic currents
can be written in the same form, for comparison:
\begin{equation}
\mathbb{V}^{t=2}_\mu =
\partial_\mu \sum_{N,\mathrm{odd}}\frac{\varkappa_N}{N+2}
\int_0^1\!du\, u^N \bar u^{N+2}\, [\mathcal{O}^V_{N}(ux)]_{l.t.}\,.
\end{equation}
Here $\mathcal{O}^V_{N}(ux)$ is the conformal operator
(\ref{eq:On}) at the space-time position $ux$.
We obtain
\begin{eqnarray}
\mathbb{V}^{t=4}_\mu&=&\frac12 \sum_{N,\text{odd}}\frac{\varkappa_N}{(N+2)^2}\int_0^1 du \,
\biggl\{(u\bar u)^{N+1}x_\mu\,[\widehat{\mathcal{O}}_N^V(u x)]_{l.t.}\,
\nonumber\\ &&{}\hspace*{2.5cm}
+\frac{N}{2} u^{N-1} \bar u^{N+2}\Big[u+\frac{1}{N+2}\Big] x^2\partial_\mu\,
[(\widehat{\mathcal{O})}_N^V(u x)]_{l.t.}\biggr\}\,,
\nonumber\\
\mathbb{A}^{t=4}_\mu&=&\frac14 \sum_{N,\text{even}}\frac{\varkappa_N N}{(N+2)^2}
\int_0^1 du \, u^{N-1} \bar u^{N+2} \Big[u+\frac{1}{N+2}\Big]
x^2\partial_\mu[\widehat{\mathcal{O}}_N^A(u x)]_{l.t.}\,,
\nonumber\\
\mathbb{X}^{t=4}&=&\frac14\sum_{N,\textrm{odd}}\frac{\varkappa_N}{(N+2)^2}
\int_0^1 du\, u^{N-1}\bar u^{N+1}\Big[1-2\frac{N+1}{N+2}\bar u\Big]
[\widehat{\mathcal{O}}_N^V(u x)]_{l.t.}\,,
\nonumber\\
\mathbb{Y}^{t=4}&=&-\frac14\sum_{N,\textrm{odd}} \frac{\varkappa_N}{(N+2)^2}
\int_0^1 du\, u^{N-1}\bar u^{N+1}\Big[1-2\frac{N+1}{N+2}\bar u
+ 2\frac{N+1}{N+3}\bar u^2\Big]
[\widehat{\mathcal{O}}_N^V(u x)]_{l.t.}\,.
\end{eqnarray}
Here $\widehat{\mathcal{O}}_N$ is defined as
the divergence of the leading-twist conformal
operator, cf. $\mathcal{O}_2$ in~Eq.\,(\ref{eq:O1O2}):
\begin{eqnarray}
\widehat{\mathcal{O}}_N(y)&=&\frac1{N+1}\frac{\partial}{\partial x^\mu}
\bigl[i\mathbf{P}^\mu,\mathcal{O}_N(y)\bigr]
=
\bigl[i\mathbf{P}^\mu,\mathcal{O}_{\mu\mu_1\ldots\mu_N}(y)\bigr] x^{\mu_1}\ldots x^{\mu_N}\,.
\end{eqnarray}
Note that the operator $\mathcal{O}_1$ in Eq.\,(\ref{eq:O1O2}), which
corresponds to $[i\mathbf{P}_\mu[i\mathbf{P}^\mu, \mathcal{O}_N]$ in our present
notation, does not contribute to the answer
for our special choice of the correlation function $T\{j_\mu(x)j_\nu(0)\}$.
The T-product with symmetric positions of the currents, $T\{j_\mu(x)j_\nu(-x)\}$,
includes both operators. The corresponding expression turns out to be much more cumbersome.
Conservation of the electromagnetic current implies that
$\partial^\mu T_{\mu\nu}(x)=0$ and
$\partial^\nu T_{\mu\nu}(x)=i[\mathbf{P}^\nu, T_{\mu\nu}(x)]$.
We have checked that these identities are satisfied up to twist-5 terms.
For completeness we give the relation for the operator $[i\mathbf{P}_\mu, \partial^\mu O(z_1,z_2)]$
entering the twist-three functions $\mathbb{V}^{t-3}_\mu$, $\mathbb{A}^{t-3}_\mu$
in terms of $\widehat{\mathcal{O}}_N$:
\begin{eqnarray}
[i\mathbf{P}_{\!\mu}, \partial^\mu O(z_1,z_2)]&=&
\frac12 S^+ \!\! \int_0^1 \!\!\!\!u du\, [i\mathbf{P}_\mu[i\mathbf{P}^\mu\!, O(uz_1,uz_2)]]
\nonumber\\&&{}+
\sum_{N}\!\varkappa_N(N\!+\!1)^2 z_{12}^N \!\int_0^1 \!\!\!dv\, v^{N}
\!\!\int_0^1 \!\!\!du\, (u\bar u)^{N+1} \widehat{\mathcal{O}}_N(v z_{12}^u x),
\end{eqnarray}
where $S^+=z_1^2\partial_{z_1}+z_2^2\partial_{z_2}+2z_1+2z_2$.
It is also possible to rewrite, v.v., all contributions of local operators
$\widehat{\mathcal{O}}_N$ in terms of the nonlocal light-ray operator
$[i\mathbf{P}_\mu, \partial^\mu O(z_1,z_2)]$, which can be advantageous in
certain applications.
\section{Typical matrix elements}
Hadronic matrix elements of the twist-4 operator $\widehat{\mathcal{O}}_N$ are of course related to those
of the leading twist, ${\mathcal{O}}_N$. For illustration, we present the corresponding explicit expressions
for the two proton states with momenta $p'\slashed{=} p$, which are relevant e.g. for virtual Compton scattering.
The leading-twist matrix elements can be parametrized as (cf.~\cite{Diehl:2003ny,Belitsky:2005qn})
\begin{align}
\langle p'|\mathcal{O}_N(n)|p\rangle =\bar u(p')\slashed{n}u(p)\sum_{k=even}^NF_{N,k}(t)\Delta_+^k P_+^{N-k}
+\frac1m\bar u(p')u(p)\sum_{k=even}^{N+1} H_{N,k}(t)\Delta_+^k P_+^{N+1-k}\,,
\end{align}
where $F_{N,k}(t)$ and $H_{N,k}(t)$ are generalized form factors corresponding to moments of the leading-twist
GPD and we used the notations $P=(p+p')/2$, $\Delta = p'-p$, $p^2=(p')^2= m^2$,
$t=\Delta^2$; $u(p)$ is the nucleon spinor.
By analogy, we define
\begin{align}
\langle p'|\widehat{\mathcal{O}}_N(n)|p\rangle =\bar u(p')\slashed{n}u(p)\sum_{k=even}^N\widehat{F}_{N,k}(t)\Delta_+^k P_+^{N-k}
+\frac1m\bar u(p')u(p)\sum_{k=even}^{N+1} \widehat{H}_{N,k}(t)\Delta_+^k P_+^{N+1-k}\,.
\end{align}
A short calculation yields
\begin{eqnarray}
\widehat F_{N,k}(t)&=&t\,F_{N,k}(t)\frac{k(2N+3-k)}{2(N+1)^2}-\left(m^2-\frac{t}4\right)F_{N,k-2}
\frac{(N-k+2)(N-k+1)}{2(N+1)^2}
\notag\\
\widehat H_{N,k}(t)&=&t\,H_{N,k}(t)\frac{k(2N+3-k)}{2(N+1)^2}-\left(m^2-\frac{t}4\right)H_{N,k-2}
\frac{(N-k+3)(N-k+2)}{2(N+1)^2}
\nonumber\\
&&{}
-m^2 \frac{(N-k+2)}{(N+1)^2} F_{N,k-2}(t)\,.
\end{eqnarray}
Note that the twist-4 matrix elements involve both finite-$t$ and target (nucleon)
mass corrections. Concrete applications will be considered elsewhere.
\section{Conclusions}
To summarize, we have given a complete expression for the time-ordered product
of two electromagnetic currents that resums all kinematic corrections
related to quark GPDs to twist-four accuracy. The results can be applied to various
two-photon processes, e.g. to the
studies of deeply-virtual Compton scattering and $\gamma^*\to (\pi,\eta,\ldots) +\gamma$
transition form factors. The twist-four terms
calculated in this work give rise to {\em both} a $\propto t/Q^2$ correction and
the target mass correction $\propto m^2/Q^2$ for DVCS,
whereas for the transition form factors these two effects
are indistinguishable as there is only one scale.
The main remaining question is whether QCD factorization itself is valid
in such reactions to twist-four accuracy, at least for kinematic contributions.
Clarification of this issue goes beyond the tasks of this study.
\section*{Acknowledgments}
The work by A.M. was supported by the DFG, grant BR2021/5-2,
and RFFI, grant 09-01-93108.
V.B. thanks Profs. A. Faessler and J. Wambach
for the invitation to the school and hospitality.
|
\section{Introduction}
An key open issue for galaxy evolution and formation models is the understanding of the different mechanisms of galaxy assembly at various cosmic epochs. In this context, the gas-phase and stellar metallicities have proven to
be important parameters to constrain the star formation history of galaxies.
A relation between galaxy mass and metallicity, first discovered by \citet{leq79} in irregular galaxies, exists for star-forming galaxies both in the nearby universe \citep{tremonti04, lam04}, and at high redshifts $z\sim0.7-3$ \citep{lam06, lam09, perez09, queyrel09, erb06, maiolino08, manucci09}. This trend for more massive galaxies
to have a higher gas-phase metallicity can be explained by various related properties: gas inflows and/or outflows, strength of the gravitational potential, efficiency of the star formation depending on stellar mass, etc. Their relative importance is currently being explored in numerical simulations.
Studying the abundance in H\,\textsc{ii} regions of nearby spiral galaxies has unveiled metallicity gradients in the local Universe \citep{pageled81, vilacostas92, considere00, pilyugin04}. In $z\sim0$ galaxies, the metallicity generally decreases from the center to the outskirts (up to $\Delta_r Z\sim-0.1$ dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${}).
The strength of these gradients seems to correlate with various parameters such as rotation velocity,
morphological type, luminosity, and the presence of a bar. Different physical processes have been proposed to explain these patterns, including radial gas-flows \citep{koeppen94}, self-regulating star formation \citep{phillipps91}, variable yields \citep{vilacostas92}, or continuous infall of gas onto the disk \citep{magrini07}.
Merging events \--- which are believed to play a substantial role in galaxy evolution \citep[eg.][]{deravel09, lopez11} \--- seem to be a key physical process in shaping the metallicity gradients of interacting galaxies \citep{rupke10}. Recent observations have suggested that galaxies involved in merging events show lower nuclear metallicities due to the infall of pristine gas into the nucleus. Merging events could account also for outliers to the mass-metallicity relation \citep{kewley06, md08, peeples09, queyrel09, alonso10, montuori10, kewley10}.
Metallicity gradients have been observed in the stellar populations of early-type galaxies as well. They appear to correlate with macroscopic properties such as, for example, stellar mass \citep[eg.][]{spolaor09}.
It has been demonstrated that a metallicity gradient in the stellar populations could survive a major merger event, although it will be weakened. These observations contrast the predictions of the simple monolithical collapse scenario \citep{white80, bekki99, koba04, dimatteo09}.
In nearby galaxies, metallicity gradients can be inferred from the observation of different emission-line ratios (in H\,\textsc{ii} regions), or from stellar absorption lines. In the distant universe ($z\gtrsim0.5$), low signal-to-noise ratio (hereafter SNR) and low spatial-resolution data make these measurements much more challenging.
However, thanks to powerful integral-field spectrographs mounted on the largest telescopes, it is now possible to determine spatially-resolved physical parameters of high redshift galaxies.
Many studies have taken advantage of this technique to study the gas dynamics of distant galaxies, from $z\sim 0.6$ \citep{flores06} up to $z\gtrsim 2$ \citep{forster09, law09, epinat09, gnerucci11}. Recently, the Mass Assembly Survey with SINFONI in VVDS \citep[MASSIV;][]{contini11} collected data of $z\sim1-2$ galaxies observed with the integral-field spectrograph SINFONI at the VLT \citep{sinfo03}. Depending on the galaxy redshift, the bright H$\alpha${} emission line is targeted in the $J$ or $H$ band. The nitrogen line [N\,\textsc{ii}]6584{} close to H$\alpha${} allows to estimate the gas-phase metallicity \emph{via} the N2 ratio ($\mathrm{N2} = \log(\mathrm{[N\,\textsc{ii}]6584}/{\mathrm{H}\alpha})$) and the corresponding abundance calibration \citep{kd02, denicolo02, pmc09}.
The aim of this paper is to investigate \---~for the first time at a redshift around 1 \--- metallicity gradients of 50 star-forming galaxies in the MASSIV sample. In order to achieve this goal, we develop a dedicated program to analyse the low SNR data cubes. Two companion papers \citep{vergani11, epinat11} discuss the associated fundamental scaling relations (e.g.~the Tully-Fischer relation for disks) and the kinematical properties of the sample galaxies.
The results of these three studies are cross-correlated in order to investigate possible relations between the metallicity gradients and global properties of galaxies at $z\sim1.2$.
The paper is organized as follows. In Section~\ref{sec:datadescr} we briefly summarise the properties of the galaxy sample, the data reduction technique, the emission-line measurements, the level of Active Galactic Nucleus (AGN) contamination, the kinematics, and the mass (both stellar and dark matter halo) estimates. In Section~\ref{sec:zgrad} we present and discuss the determined metallicity gradients. Our conclusions are drawn in Section~\ref{sec:conclu}.
Throughout the paper, we assume a $\Lambda$CDM cosmology with $\Omega_m=0.3$, $\Omega_\Lambda=0.7$ and $H_0 = 70~ \mathrm{km s}^{-1}\mathrm{ Mpc}^{-1} $.
\section{Data Description}
\label{sec:datadescr}
\subsection{Target selection, observations and data reduction}
The galaxy sample studied in this paper is the ``first epoch'' sample of the MASSIV project (ESO Large Program, PI.: T. Contini). A full description of the sample can be found in \cite{contini11}.
We briefly summarise some properties of this sample of 50 galaxies below.
The galaxies were selected from the VIMOS VLT Deep Survey (VVDS) in the deep \citep[$I_{AB}<24$;][]{lefevre05}, ultra-deep
($I_{AB}<24.75$; Le F\`evre et al., in prep ) $\mathrm{RA} = 02\mathrm{h}$, and wide \citep[$I_{AB}<22.5$;][]{garilli08} $\mathrm{RA}=14\mathrm{h}, 22\mathrm{h}$ fields. The galaxies were chosen according to their redshift such that their H$\alpha${} line (or [O\,\textsc{iii}]5007 in a few cases) was visible in the $J$ or $H$ band, and was not affected by a bright OH sky-line. Galaxies were selected to be star-forming on the basis of their [O\,\textsc{ii}]3727 emission line strength. The observations have been performed between April 2007 and August 2009. The general properties (RA, DEC, redshift, $I$-band magnitude) and characteristic observational configuration (spectral band, adaptive optics or not, exposure time, spatial resolution) of each galaxy are gathered in Table~\ref{gene}. Most ($85$\%) of the galaxies in this ``first epoch'' sample have been observed in a seeing-limited mode (with a spatial sampling of $0.125$\arcsec{}). However, seven galaxies have been acquired with Adaptive Optics assisted with a Laser Guide Star (AO/LGS, $0.05$\arcsec{} spatial sampling). Among the 50 ``first epoch" MASSIV galaxies observed with SINFONI, four galaxies have not been detected \citep[see][for details]{epinat11}.
The data reduction was performed with the ESO SINFONI pipeline, using the standard master calibration files provided by ESO. The absolute astrometry for the SINFONI data cubes was derived from nearby bright stars also used for PSF measurements. Custom \textsf{IDL} and \textsf{Python} scripts have been used to flux calibrate, align, and combine all the individual exposures. For each galaxy a non sky-subtracted cube was also created, mainly to estimate the effective spectral resolution. For more details on data reduction, we refer to \cite{epinat11}.
$I$-band images for all the galaxies were obtained through CFHT Megacam imaging (for the 22h and
02h field, from the CFHTLS ``best seeing'') and CFHT-12K imaging (for the 14h field). These images were used for two purposes: (i) refinement of the SINFONI astrometry, using the relative position of the PSF star, and (ii), deriving morphological parameters \citep[used as inputs to the kinematics modeling, see][]{epinat11}.
\begin{table*}
\caption{General properties and SINFONI observing log of the 50 ``first
epoch'' galaxies of MASSIV. Column~1: identifier in the VVDS catalog,
Columns~2 \& 3: RA \& DEC in
degrees, Column~4: spectral band of the observation, Column~5:
seeing-limited mode (0) or Adaptive Optics (1), Column~6: on-source
total exposure time, Column~7: $I$-band magnitude in the AB system,
Column~8: Mean spatial resolution (PSF) as measured on a nearby
star, Column~9: redshift from H$\alpha${} (or [O\,\textsc{iii}]5007 when marked with
an asterisk) as measured with SINFONI. For the four galaxies (VVDS020126402,
VVDS020217890, VVDS020306817, and VVDS220071601) which have not been
detected with SINFONI, redshift comes from VIMOS spectra}\label{gene}
\begin{center}
\begin{tabular}{ccccccccc}
\hline
\hline
Galaxy & RA (J2000) & Dec (J2000) & Band & AO/LGS &$t_\mathrm{exp}$ & $I_\mathrm{AB}$ & Mean PSF & $z_\textrm
{SINFONI}$ \\
& [deg] & [deg] & & & [sec] & [mag] & [\arcsec] & \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) \\
\hline
020106882 & 36.340833 & -4.771834 & H & 0 & 4800 & 23.18 & 0.49 & 1.3991 \\
020116027 & 36.463055 & -4.751243 & H & 0 & 4500 & 22.88 & 0.60 & 1.5302 \\
020126402 & 36.298576 & -4.727812 & J & 1 & 3600 & 22.99 & $\dots$ & 1.2332 \\
020147106 & 36.689110 & -4.679830 & H & 0 & 7200 & 22.51 & 0.65 & 1.5195 \\
020149061 & 36.771769 & -4.674816 & H & 0 & 4800 & 22.56 & 0.85 & 1.2905 \\
020164388 & 36.712244 & -4.639131 & H & 0 & 4800 & 22.51 & 0.83 & 1.3547 \\
020167131 & 36.697114 & -4.632054 & J & 0 & 7200 & 23.01 & 0.68 & 1.2246$^*$ \\
020182331 & 36.684314 & -4.597755 & H & 0 & 10800 & 22.73 & 0.74 & 1.2290 \\
020193070 & 36.327984 & -4.572212 & J & 0 & 7200 & 23.41 & 0.58 & 1.0279 \\
020208482 & 36.319735 & -4.536649 & J & 0 & 7200 & 23.25 & 0.58 & 1.0375 \\
020214655 & 36.597688 & -4.523057 & J & 0 & 4800 & 23.05 & 0.87 & 1.0395 \\
020217890 & 36.613174 & -4.514397 & H & 0 & 7200 & 23.99 & $\dots$ & 1.5129 \\
020239133 & 36.679087 & -4.475228 & J & 0 & 4800 & 22.85 & 0.79 & 1.0194 \\
020240675 & 36.725581 & -4.471541 & H & 0 & 4800 & 23.45 & 0.85 & 1.3270 \\
020255799 & 36.691075 & -4.437713 & J & 0 & 4800 & 23.63 & 0.76 & 1.0351 \\
020261328 & 36.796038 & -4.425444 & H & 0 & 3600 & 23.90 & 0.62 & 1.5290 \\
020278667 & 36.492598 & -4.386738 & J & 0 & 7200 & 23.25 & 0.65 & 1.0516 \\
020283083 & 36.628634 & -4.376830 & H & 0 & 4800 & 23.07 & 0.78 & 1.2818 \\
020283830 & 36.620565 & -4.375444 & H & 0 &7200 & 22.92 & 0.77 & 1.3949 \\
020294045 & 36.446413 & -4.352110 & J & 0 & 7200 & 22.80 & 0.59 & 1.0028 \\
020306817 & 36.459649 & -4.323037 & J & 0 & 7200 & 23.29 & $\dots$ & 1.2225 \\
020363717 & 36.598723 & -4.199505 & H & 0 & 4800 & 22.61 & 0.64 & 1.3339 \\
020370467 & 36.561244 & -4.184841 & H & 0 & 4800 & 23.46 & 0.71 & 1.3338 \\
020386743 & 36.808285 & -4.149879 & J & 0 & 7200 & 22.58 & 0.73 & 1.0487 \\
020461235 & 36.696290 & -4.398832 & J & 0 & 7200 & 22.64 & 0.63 & 1.0349 \\
020461893 & 36.801068 & -4.386450 & J & 0 & 4800 & 23.45 & 0.60 & 1.0486 \\
020465775 & 36.747379 & -4.316665 & H & 0 & 4800 & 23.15 & 0.88 & 1.3583 \\
140083410 & 209.460955 & 4.294251 & J & 0 & 4800 & 21.82 & 0.69 & 0.9435 \\
140096645 & 209.609695 & 4.329940 & J & 0 & 7200 & 22.28 & 0.56 & 0.9655 \\
140123568 & 208.990111 & 4.405582 & J & 0 & 7200 & 23.43 & 0.76 & 1.0012 \\
140137235 & 209.053257 & 4.442166 & J & 0 & 4800 & 22.38 & 0.76 & 1.0445 \\
140217425 & 209.485039 & 4.643635 & J & 0 & 6000 & 21.58 & 0.95 & 0.9792 \\
140258511 & 210.081944 & 4.746065 & H & 0 & 4800 & 21.17 & 0.49 & 1.2423 \\
140262766 & 209.981154 & 4.758375 & H & 0 & 7200 & 23.68 & 0.51 & 1.2836 \\
140545062 & 209.898275 & 5.508636 & J & 0 & 7200 & 22.42 & 0.70 & 1.0408 \\
220014252 & 334.440374 & 0.477630 & H & 0 & 7200 & 22.04 & 0.70 & 1.3105 \\
220015726 & 333.926894 & 0.484332 & H & 0 & 7200 & 22.42 & 0.46 & 1.2933 \\
220071601 & 334.506537 & 0.759637 & H & 1 & 4800 & 21.74 & $\dots$ & 1.3538 \\
220148046 & 333.657514 & 1.139144 & H & 1 & 4800 & 22.39 & 0.27 & 2.2441$^*$ \\
220376206 & 335.024084 & -0.139356 & H & 0 & 7200 & 21.78 & 0.50 & 1.2445 \\
220386469 & 334.985762 & -0.050878 & J & 1 & 2400 & 22.10 & 0.23 & 1.0226 \\
220397579 & 335.152164 & 0.029634 & J & 0 & 7200 & 22.42 & 0.64 & 1.0379 \\
220544103 & 333.857118 & 0.110982 & H & 0 & 7200 & 22.38 & 0.76 & 1.3973 \\
220544394 & 333.600634 & 0.113027 & J & 0 & 7200 & 22.15 & 0.58 & 1.0101 \\
220576226 & 334.047658 & 0.275121 & J & 0 & 7200 & 21.80 & 0.58 & 1.0217 \\
220578040 & 334.267094 & 0.282327 & J & 0 & 7200 & 22.30 & 0.62 & 1.0462 \\
220584167 & 333.845992 & 0.313059 & H & 0 & 7200 & 21.96 & 0.75 & 1.4655 \\
220596913 & 333.621601 & 0.371914 & H & 1 & 7200 & 21.80 & 0.18 & 1.2658 \\
910193711 & 36.442825 & -4.542796 & H & 1 & 4800 & 22.69 & 0.27 & 1.5564 \\
910279515 & 36.401024 & -4.354381 & H & 1 & 4800 & 23.71 & 0.21 & 1.4013 \\
\hline
\end{tabular}
\end{center}
\end{table*}
\subsection{Measurement of emission lines}\label{el}
As mentioned in the previous section, galaxies were selected to be star-forming such that emission lines from H\,\textsc{ii} regions are visible in the near infrared. Given the redshift range of MASSIV galaxies, we targeted, in most of the cases, the rest-frame optical H$\alpha${} emission line in the $J$ (for $z <1.2$) or $H$ (for $z>1.2$) band. For four objects (VVDS020126402, 020167131, 020306817, and 220148046) we observed the [O\,\textsc{iii}]5007 line which allowed to increase the redshift window (two of them were not detected). These latter galaxies are not considered in this paper as we cannot derive any metallicity from the [O\,\textsc{iii}]5007 emission-line alone. In the former cases, we took advantage of the proximity of the nitrogen line to H$\alpha${} to derive the metallicity of a majority of the galaxies using the [N\,\textsc{ii}]6584{}/H$\alpha${} line ratio as a proxy for oxygen abundance (see sect.~\ref{sec:zgrad}).
\subsubsection{Integrated fluxes}\label{intfl}
The global integrated flux of each emission line was calculated using a mask designed for each galaxy. The flux was collected over a region where the signal-to-noise ratio of the H$\alpha${} line is above a fixed threshold of $SNR>2$. The SNR maps were produced using a Gaussian-fitting procedure written in \textsf{IDL}, already used and described in \cite{epinat11}. A Gaussian spatial smoothing of $2\times 2$ spaxels (0.25\arcsec$\times$ 0.25\arcsec) was used, increasing the SNR without degrading the final spatial resolution ($\sim 0.65$\arcsec\ on average). The line position map produced by this procedure was used to shift in wavelength every single spectrum such as to compensate for the Doppler effect corresponding to the measured H$\alpha${} velocity at that position. In this way, the broadening of the line by large-scale rotation in the spatially-integrated spectrum was largely removed and the line profile more easily fitted with a Gaussian, increasing the accuracy of line measurements.
The H$\alpha${} fluxes were measured on flux-calibrated integrated spectra using a Gaussian fit and a flat continuum. These fluxes have been corrected
for dust reddening using the extinction coefficient derived from the SED fitting \citep[see][]{contini11, vergani11}. The line ratios, on the other hand, were determined from the integrated spectra in counts. Indeed, the flux calibration procedure would have added noise on top of the already low SNR data. Further, the lines of interest (namely, H$\alpha${} and [N\,\textsc{ii}]6584) are close enough in wavelength to assume that the sensitivity curve is constant over the wavelength range of interest. The same argument justifies the fact that we did not correct the emission-line ratios for differential extinction. Table~\ref{eml} lists our final integrated H$\alpha${} flux, N2 emission line ratio
($N2 = \log(\mathrm{[N\,\textsc{ii}]6584{}}/\mathrm{H\alpha})$ and star formation rates (SFR, corrected or not for dust reddening). Among the 44 galaxies detected in H$\alpha${} with SINFONI, we have been able to measure the [N\,\textsc{ii}]6584{} emission-line in 34 galaxies and hence derive integrated N2 emission-line ratio and metallicity for these objects.
\tabcolsep1.40mm
\begin{table}[!ht]
\caption{H$\alpha${} flux, $N2= \log(\mathrm{[N\,\textsc{ii}]6584{}}/\mathrm{H\alpha}$) emission-line ratio, and H$\alpha${}-based star formation rates
(corrected \--- and not \--- for dust reddening) for each galaxies in our sample.}\label{eml}
\centering
\begin{tabular}{crrrr}
\hline
\hline
Galaxy & F(H$\alpha$) & N2 & $SFR_{\rm{H}\alpha}$ & $SFR_{\rm{H}\alpha}^{\rm cor}$ \\
& [${10}^{-17}$ergs$^{-1}$} \newcommand{\car}{$^{-2}${}cm\car{}] & & [M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}${} yr$^{-1}$} \newcommand{\car}{$^{-2}${}] & [M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}${} yr$^{-1}$} \newcommand{\car}{$^{-2}${}] \\
(1) & (2) & (3) & (4) & (5) \\
\hline
020106882 & $14.0 \pm 0.8$ & $-0.71 \pm 0.12$ & 13.3 & 38.1 \\
020116027 & $12.4 \pm 2.9$ & $-1.02 \pm 0.15$ & 14.7 & 42.2 \\
020126402 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
020147106 & $46.8 \pm 5.8$ & $-0.95 \pm 0.14$ & 54.3 & 92.1 \\
020149061 & $20.6 \pm 1.6$ & $\dots$ & 16.0 & 45.9 \\
020164388 & $24.1 \pm 1.3$ & $-0.86 \pm 0.10$ & 21.1 & 46.5 \\
020167131 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
020182331 & $15.2 \pm 5.0$ & $\dots$ & 10.5 & 51.1 \\
020193070 & $ 6.7 \pm 0.9$ & $-0.56 \pm 0.13$ & 3.0 & 14.4 \\
020208482 & $ 2.6 \pm 0.5$ & $\dots$ & 1.2 & 2.6 \\
020214655 & $23.3 \pm 1.1$ & $-0.68 \pm 0.06$ & 10.6 & 51.7 \\
020217890 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
020239133 & $ 7.8 \pm 1.0$ & $\dots$ & 3.4 & 16.5 \\
020240675 & $ 8.5 \pm 1.0$ & $\dots$ & 7.1 & 15.7 \\
020255799 & $ 5.8 \pm 0.7$ & $-0.65 \pm 0.15$ & 2.6 & 12.8 \\
020261328 & $ 9.5 \pm 0.9$ & $-1.06 \pm 0.48$ & 11.2 & 19.0 \\
020278667 & $ 2.4 \pm 1.3$ & $\dots$ & 1.1 & 4.2 \\
020283083 & $13.4 \pm 1.0$ & $-0.70 \pm 0.12$ & 10.3 & 17.4 \\
020283830 & $11.0 \pm 1.1$ & $-0.82 \pm 0.17$ & 10.3 & 50.4 \\
020294045 & $15.5 \pm 2.8$ & $\dots$ & 6.5 & 14.3 \\
020306817 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
020363717 & $37.4 \pm 4.3$ & $-0.87 \pm 0.09$ & 31.5 & 53.4 \\
020370467 & $17.3 \pm 2.9$ & $-0.69 \pm 0.11$ & 14.6 & 71.2 \\
020386743 & $17.4 \pm 1.7$ & $-0.91 \pm 0.08$ & 8.1 & 39.4 \\
020461235 & $ 7.5 \pm 2.0$ & $-0.67 \pm 0.12$ & 3.3 & 9.6 \\
020461893 & $14.5 \pm 1.2$ & $\dots$ & 6.7 & 14.9 \\
020465775 & $14.5 \pm 2.8$ & $\dots$ & 12.8 & 62.2 \\
140083410 & $47.0 \pm 3.3$ & $-0.72 \pm 0.08$ & 16.8 & 37.1 \\
140096645 & $55.4 \pm 4.2$ & $-0.25 \pm 0.02$ & 20.9 & 102.1 \\
140123568 & $ 5.8 \pm 1.2$ & $-0.69 \pm 0.10$ & 2.4 & 11.8 \\
140137235 & $ 9.6 \pm 2.3$ & $-0.79 \pm 0.13$ & 4.4 & 21.6 \\
140217425 & $104.7 \pm 4.9$ & $-0.44 \pm 0.02$ & 41.0 & 200.0 \\
140258511 & $42.5 \pm 2.7$ & $-0.51 \pm 0.10$ & 30.0 & 146.4 \\
140262766 & $13.1 \pm 5.0$ & $-0.75 \pm 0.15$ & 10.0 & 10.0 \\
140545062 & $29.4 \pm 3.3$ & $-0.75 \pm 0.06$ & 13.4 & 22.7 \\
220014252 & $50.3 \pm 3.1$ & $-0.67 \pm 0.08$ & 40.5 & 197.5 \\
220015726 & $47.5 \pm 6.7$ & $-0.69 \pm 0.04$ & 37.0 & 106.4 \\
220071601 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
220148046 & $\dots$ & $\dots$ & $\dots$ & $\dots$ \\
220376206 & $72.4 \pm 5.1$ & $-1.04 \pm 0.10$ & 51.2 & 249.7 \\
220386469 & $17.1 \pm 2.4$ & $-1.14 \pm 0.16$ & 7.5 & 36.5 \\
220397579 & $65.1 \pm 9.1$ & $-1.07 \pm 0.07$ & 29.4 & 143.2 \\
220544103 & $55.0 \pm 1.9$ & $-0.56 \pm 0.09$ & 52.2 & 117.5 \\
220544394 & $24.3 \pm 1.7$ & $-0.89 \pm 0.07$ & 10.3 & 50.1 \\
220576226 & $31.2 \pm 1.6$ & $-0.69 \pm 0.04$ & 13.6 & 66.3 \\
220578040 & $20.5 \pm 1.9$ & $-0.80 \pm 0.07$ & 9.4 & 20.8 \\
220584167 & $64.3 \pm 1.3$ & $-0.91 \pm 0.10$ & 68.6 & 202.6 \\
220596913 & $30.0 \pm 1.0$ & $\dots$ & 23.2 & 36.1 \\
910193711 & $32.9 \pm 18.4$ & $-0.68 \pm 0.09$ & 40.6 & 197.7 \\
910279515 & $14.9 \pm 2.1$ & $-0.83 \pm 0.21$ & 14.2 & 69.0 \\
\hline
\end{tabular}
\end{table}
\subsubsection{Spatially-resolved flux and line ratio}
For many galaxies in our sample, the low SNR of the data cubes prevents us from measuring in each spaxel, with high confidence, the emission lines surrounding H$\alpha${} : namely [N\,\textsc{ii}]6584, [S\,\textsc{ii}]6717,31{}, and [O\,\textsc{i}]6300 which is detected only in one galaxy: VVDS140096645.
These lines might either not be bright enough to be detected, or a bright OH sky-line is too close. To
increase the SNR of our final spectra, we summed the spectra of the spaxels gathered in a specified region, and measured the line fluxes in the resulting spectra.
We developed a specific \textsf{Python} procedure to do this analysis within our data cubes.
The program allowed to define a region spaxel-by-spaxel, or to define a region with contours on 2D maps (such as H$\alpha${} flux, SNR, etc maps). In a given region, the spaxel-to-spaxel line shift due to the rotation velocity of the gas along the line-of-sight was fully corrected for (see \S\ref{intfl}). The integrated spectrum in a defined region was then fitted with a flat continuum and two Gaussians, one for H$\alpha${} and one for [N\,\textsc{ii}]6584{} (the sulfur doublet was neglected for this analysis as it was detected in a few cases only). The width of each Gaussian was set to be the same as the emission of the collisional and recombination lines traces the same ionised gas in the galaxies. The
fit was weighted using the corresponding non sky-subtracted spectrum, i.e.~giving a lower weight to the channels affected by the subtraction of a strong sky-line. We then estimated the $1\sigma$ error on the fluxes with a Monte-Carlo technique, in which the fit was considered to be a ``noise-free'' model, to which we added a Gaussian noise with a standard deviation corresponding to the residuals from the initial spectrum minus the ``noise-free'' model. This operation was repeated a hundred of times, which gave a set of parameters on which the $1\sigma$ deviation was computed.
\fig{prog} shows a result of our fits in different regions for two galaxies.
Finally, we estimate spatially-resolved H$\alpha$\ and [N\,\textsc{ii}]6584\ emission-line flux for 26/34 ($\sim 75$\%) galaxies in the sample. The remaining 8 galaxies have either a too low SNR or
the [N\,\textsc{ii}]6584\ emission-line is polluted by sky-line residuals.
\begin{figure}[t]
\centering
\raisebox{0.7cm}{\includegraphics[width=0.5\linewidth]{220578040_contours.pdf}}\includegraphics[width=0.5\linewidth]{220578040_sp.pdf}
\includegraphics[width=0.5\linewidth]{140217425_sp.pdf}\raisebox{0.7cm}{\includegraphics[width=0.5\linewidth]{140217425_contours.pdf}}
\caption{Examples
of spatially-resolved emission-lines measurements in two MASSIV
galaxies (Top: VVDS220578040, bottom: VVDS140217425). Color-coded
images show the regions defined with H$\alpha${} contours used to derive
the metallicity gradients. The 1D spectra integrated over the
different regions are shown with solid black lines. Model (in red)
and sky (dashed line) spectra are also shown. The sky spectrum is
used as a weight in the fitting procedure. Region 1 is the outer most with the region number increasing towards the center.}\label{prog}
\end{figure}
\subsection{AGN contamination}
A recent study \citep{wright10} has shown that unveiling the presence of AGN in high-redshift galaxies is a difficult exercise. In the case of metallicity studies, in which abundances are deduced from the ratio of
different emission lines of the ionised gas, it is critical to check that the intensity and width of these lines are due to star formation and not related to any non-thermal nuclear activity. The common way to disentangle AGN contribution from star-forming galaxies consists in comparing the relative intensity of the main nebular
emission lines (mainly [O\,\textsc{iii}]5007, H$\beta$, H$\alpha$, and [N\,\textsc{ii}]6584{}) in a diagnostic diagram, so-called BPT diagram \citep{bpt81, kewley01}. Various physical conditions in the ISM \--- SFR, ionisation parameter, metallicity and/or chemical composition \--- have been invoked to explain the fact that some high-$z$ star-forming galaxies lie in the transition region of the local BPT diagram (as defined by the SDSS galaxies) between
star-forming galaxies and AGN hosts. \citet{wright10} have been able \--- thanks to high resolution adaptive optics observations \--- to subtract the active nuclear emission in a $z\sim 1.6$ galaxy (HDF-BMZ1299), and have shown that the residual extended star-forming emission was characteristic of a local SDSS star-forming galaxy, whereas the integrated emission would have placed the object in the transition region.
\begin{figure}[!ht]
\centering
\includegraphics[width=0.5\linewidth]{histo_N2.pdf}\includegraphics[width=0.5\linewidth]{c_vs_n.pdf}
\caption{{\it Left}: Distribution of the N2 ratio in the MASSIV ``first
epoch'' sample. The black line represents the ratio integrated over
the whole galaxy. The red line is the N2 ratio measured in the
nuclear region. {\it Right}: N2 global ratio versus N2 nuclear
ratio.}\label{distrib}
\end{figure}
The nature of our observations did not give us simultaneous access to the set of emission lines
([O\,\textsc{iii}]5007, H$\beta$, H$\alpha$, and [N\,\textsc{ii}]6584 or [S\,\textsc{ii}]6717,31{}) commonly used in standard diagnostic diagrams. However, for all but two objects in our sample, the emission-line ratio $\mathrm{N2} =
\log(\textrm[N\,\textsc{ii}]6584{}/\textrmH$\alpha$)$ is lower than $-0.5$, with a median $\mathrm{N2}$ value of $-0.72$. Such low values are indicative for a very low contamination by AGN in our sample \citep[eg.][]{bpt81}.
For 24 galaxies of our sample we calculated, following \citet{wright10}, the N2 ``concentrated ratio'', corresponding to the value in the nuclear region of the galaxy. We defined the nuclear region as the spaxel with the highest H$\alpha${} flux along with its 8 nearest neighbours (corresponding to a $0.7''$ diameter aperture, matching our mean spatial resolution of $0.65''$). We assumed this aperture to be small enough to probe the inner
nucleus part as objects usually span from $1''$ to $2''$ with $SNR>2$ in our observations.
Fig.~\ref{distrib} shows i) the distribution of the 24 galaxies as a function of their global and nuclear N2 ratio (left panel) and ii) the relation between the global and nuclear N2 ratios for each galaxy (right panel). The median values of N2 for each distribution are not very different ($\Delta\sim -0.07$). The median nuclear N2 ratio is lower than the global ratio which would not be the case if a significant fraction of our sample galaxies were hosting an AGN. When comparing the global ratio to the nucleus ratio distribution, the highest bin does not shift and contains a single object . We investigated in more detail the galaxy in this bin: VVDS140096645. It shows the following high N2 ratios: $N2_\textrm{global}=-0.252$ and $N2_\textrm{nuclear}=-0.294$. Looking further into its integrated spectrum (global and nuclear, see \fig{645}), we noticed that: (i) the emission lines are broad, which is a possible sign of nuclear activity, (ii) the two nitrogen lines are clearly visible, as is the sulfur doublet, and the [O\,\textsc{i}]6300 line, which altogether are characteristics of LINER galaxies, often associated to AGNs \citep{heck80} \--- or violent episodes of star-formation in high metallicity galaxies \citep{terl85}.
\begin{figure}[t]
\centering
\includegraphics[width=\linewidth]{645.pdf}
\caption{Integrated spectra of VVDS140096645. {\it Top}: spectrum of the
disk (total integrated with central contribution removed). {\it Middle}:
spectrum of the central region surrounding the peak of H$\alpha${}
flux. {\it Bottom}: spectrum from the whole spatial extent. From left to
right, the following emission lines appear: [\textsc{O\,i}]6300,
[\textsc{N\,ii}]6548, H$\alpha${}, [\textsc{N\,ii}]6584,
[\textsc{S\,ii}]6717, [\textsc{S\,ii}]6731. The red and blue lines
are least-square fits to the data.}\label{645}
\end{figure}
The galaxy with the next highest global N2 value is VVDS140258511
($N2_\textrm{global}=-0.509$ and $N2_\textrm{nuclear}=-0.345$). Its global and nuclear spectra show no obvious features of nuclear activity, and the N2 ratio are both $< -0.3$.
We conclude that our sample does not suffer from significant AGN contamination. The only candidate for nuclear activity is VVDS140096645. Similar conclusions were reached based on composite 1D VIMOS and SINFONI spectra of the MASSIV sample \citep[see][]{contini11}.
\subsection{Kinematics and close environment classification}\label{kin}
A kinematics and close environment classification of the MASSIV galaxies has been performed and is described in details in \cite{epinat11}. This classification is based on the $I$-band morphology, the H$\alpha$ flux maps,
the H$\alpha$ velocity fields and their modeling assuming a rotating disk.
In the present paper, we exploit two types of classes: a) the dynamical state of galaxies (rotating or non-rotating),
and b) the close environment of galaxies. The first class relies i) on the agreement between the position angles of the major axis deduced from the morphology and from the kinematics, and ii) on the accuracy of the rotating disk model.
In the case the average SNR is lower than 5, this classification is believed to be not reliable and is thus not used. The second class relies on the detection of companions in both $I$-band image and H$\alpha$
maps at a similar redshift than the main source. In some cases, the kinematics maps (velocity fields and velocity dispersion maps) suggest the presence of some companions about merging or along the line-of-sight.
A confidence flag was assigned to the environment classification, ranging from A (>90\%, confident) to C ($\sim 50$\%).
We used these classifications to help us constraining the origin of the metallicity gradients in our MASSIV star-forming galaxies.
\subsection{Stellar and dark matter halo masses}
\label{sec:masses}
The stellar masses used in our study were obtained using the SED fitting technique. Stellar population synthesis models \citep{BC03} were used to match photometric and spectroscopic data from our MASSIV sample, using
the GOSSIP tool \citep{franzetti08}. A \citet{salpeter55} IMF was assumed, stellar formation timescales and the extinction parameter $E(B-V)$ were allowed to range from $0.1$ to $15$~Gyr, and from $0$ to $0.3$, respectively.
The GOSSIP tool returned the best-fit parameters as well as a Probability Distribution Function (PDF) for each of them, following \citet{walcher08}. The median and the standard deviation of the PDF were used to recover the parameter
estimates and their associated errors. These latter values were used for this study. The procedure is described in detail in \cite{contini11}.
The dark matter (DM) halo masses ($M_{\rm halo}$) used hereafter were computed using a spherical virialized collapse model \citep{peebles80, whiteFrenck91, mowhite02}:
\begin{equation}
\label{eq:mhalo}
M_{\rm halo} = 0.1
H_0^{-1}\mathcal{G}^{-1}\Omega_m^{-0.5}(1+z)^{1.5}V_{\rm max}^3
\end{equation}
where $\mathcal{G}$ is the universal gravitational constant, $z$ is the redshift of the galaxy and $V_{\rm max}$ is the maximum rotational velocity computed in \cite{epinat11}. In order to compute associated uncertainties,
the Monte Carlo method used in \cite{epinat11} for the uncertainties on $V_{\rm max}$ was extended until the halo mass. The contributions of the two sources of uncertainty on the velocity (inclination and modeling) were added
quadratically to compute the final uncertainty.
Equation \ref{eq:mhalo} makes the assumption that the plateau has been reached and that $V_{\rm max}$ traces the halo circular velocity. However, this assumption is probably not correct for non-rotating systems.
The stellar and DM halo masses derived for our galaxies are listed in Table~\ref{metal}.
\section{Metallicity gradients}
\label{sec:zgrad}
\begin{figure}[t]
\centering
\includegraphics[width=\linewidth]{N2_comp_paper.pdf}
\caption{Comparison between different calibrations used to derive the oxygen abundance from the N2 parameter. All these calibrations
are consistent within their intrinsic dispersion ($\sim 0.3-0.4$ dex), especially in the range of N2 parameter ($-1.0 < {\rm N2} < -0.5$) covered by MASSIV galaxies (see Table~\ref{eml} and Figure~\ref{distrib}).}\label{compacalib}
\end{figure}
In the local Universe metallicity gradients in spiral galaxies are commonly measured \citep{pageled81, vilacostas92, considere00, pilyugin04}. These gradients are generally negative (metallicity decreasing from the center to the outer parts) and their amplitude is typically $\gtrsim-0.1$~dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${} as traced by the metallicity in H\,\textsc{ii} regions. In the Milky Way, a gradient of $-0.07$~dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${} is observed.
Several physical processes can be responsible for such gradients \citep{gotz92}. Radial gas flows draining metal-rich interstellar gas from the outer parts of the galaxy into the center are believed to play a key role \citep{tinsley78}. It requires that the infall timescale of gas onto the disk is faster than the star formation
timescale. Indeed, \citet{koeppen94} has shown that the presence of both infall of pristine gas and radial flows into the disk is very efficient in creating an abundance gradient. \citet{phillipps91} further suggested that self-regulating star formation rates varying with the galactocentric distance could generate such gradients.
In contrast, for the case of interacting galaxies, \cite{rupke10} claim that radial flows of low-metallicity gas from the merging galaxies can explain the low oxygen abundances observed at their center. In these cases, the radial mixing of gas may flatten existing metallicity gradients.
At high redshifts, metallicity gradients are harder to detect as the gas-phase metallicity of individual galaxies can only be measured with a limited accuracy. Collisional emission lines from the metals and recombination lines from hydrogen and helium are the only indicators allowing to estimate the oxygen abundance in the gas-phase, needed for a direct comparison with the local galaxies.
In the present study, we used the ratio of the [N\,\textsc{ii}]6584 nitrogen line to the H$\alpha${} Balmer line as a proxy for the oxygen abundance. Computing this particular ratio has two practical advantages: the [N\,\textsc{ii}]6584{} and H$\alpha${} emission lines are very close in wavelength, so that {\it i)} the differential extinction due to dust attenuation can be neglected; and {\it ii)} the relative flux calibration can be considered to be constant over the spectral range.
To derive from this ratio the gas-phase oxygen abundance, we used the calibration proposed by \citet{pmc09}:
\begin{equation}
12 + \log(O/H) = 9.07 + 0.79\times \mathrm{N2}, \textrm{where } \mathrm{N2}=\log \frac{\mathrm{[N\,\textsc{ii}]6584{}}}
{\mathrm{H}\alpha}
\end{equation}
This calibration was computed using emission-line objects (star-forming galaxies and H\,\textsc{ii} regions) in the nearby universe with an accurate oxygen abundance obtained from the measurement of the electronic temperature. The calibration has an intrinsic scatter of 0.34 dex, mainly due to the second-order dependence of the N2 parameter on the ionisation parameter and on the nitrogen-to-oxygen abundance ratio. As shown in Figure~\ref{compacalib}, the PMC09 calibration of the N2 parameter used in this analysis is totally consistent, for both the low- and high-metallicity regime,
with other calibrations of the same parameter found in the literature. Although the N2 parameter was calibrated by PMC09 using data with a determination of O/H abundance based only on the ``T-method", and hence not very well defined for the high-metallicity regime, the PCM09 calibration agrees very well within its dispersion limits (about $0.3-0.4$ dex) with other relations that are partially based on photoionization models or strong-line
determinations of the metallicity \citep{denicolo02, PP04, nagaoetal06}. This is especially true in the range of N2 parameter ($-1.0 < {\rm N2} < -0.5$) covered by MASSIV galaxies (see Table~\ref{eml} and Figure~\ref{distrib}).
\subsection{Observed metallicity gradients}
\begin{figure*}[t]
\centering
\includegraphics[width=\linewidth]{gradz_r.pdf}
\caption{Metallicity gradient for the 26 MASSIV galaxies with
spatially-resolved metallicities. The $x$-axis represents the mean
radius (in kpc) of the region relative to the H$\alpha$
center. The $y$-axis indicates the corresponding metallicity in
dex ($12+\log(O/H)$). The red lines are the best fits to the data,
taking into account the errors on the metallicities. The yellow/blue/green
regions represent the $1\sigma$ errors associated to the
gradients. Blue label indicates the galaxies for which the
gradient is positive within $1\sigma$, the green ones are those
for which it is negative within the same limits. For each galaxy
we have indicated the dynamical (Rot=rotating disk, NRot=no rotation)
and environment (Iso=isolated, Int=interacting) classes.}\label{grad}
\end{figure*}
As reported in \S\ref{el}, we measured H$\alpha$ and [N\,\textsc{ii}]6584{} line fluxes in spatially-resolved regions for 26 galaxies. These regions, defined by H$\alpha${} isoflux contours, are centered around the H$\alpha$ peak in each galaxy, which most often corresponds to the kinematical center of the galaxy. The widths of the annular regions were adjusted such as to have a high enough SNR for the measurement of the [N\,\textsc{ii}]6584{} line and,
consequently, the N2 ratio. This allowed to quantify the radial behaviour of the H$\alpha${} and [N\,\textsc{ii}]6584{} lines, i.e.~of the gas-phase metallicity from the inner to the outer parts of each galaxy.
The derived metallicity gradients and the integrated metallicities are listed in Tab~\ref{metal}. The metallicity is estimated as a
function of radius --- each region having a mean radius and a metallicity (uncertainties on metallicity estimates are dominated by measurements errors). The radius of a region corresponds to the mean radius between its outer and inner contours. The radius of a contour is approximated by the radius of a perfect circle with the same perimeter. Depending on the galaxy inclination, the intrinsic deprojected radius can be underestimated up to a factor of $\sim 1.5$. Metallicities of the different regions in each galaxies are plotted in
\fig{grad}, along with the best fit line to the data. The figure shows that the detected radial gradients are in general very weak, some being positives and other negatives.
\begin{table}[!ht]
\caption{Integrated metallicity ($Z = 12 + \log(O/H)$), metallicity gradient, stellar and dark matter halo masses
of the MASSIV ``first epoch'' sample galaxies.}\label{metal}
\centering
\begin{tabular}{crrrr}
\hline
\hline
Galaxy & $Z$ & $\Delta_r Z$ & $\log(M^\star)$ & $\log(M_{\rm h})$ \\
& & [dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${}] & [$M_\odot$] & [$M_\odot$] \\
(1) & (2) & (3) & (4) & (5) \\
\hline
020106882 & $8.73$ & $0.037 \pm 0.026$ & $9.99_{-0.22}^{+0.22}$ & $11.59_{-0.38}^{+0.20}$\\
020116027 & $8.42$ & $0.03 \pm 0.024$ & $10.09_{-0.23}^{+0.23}$ & $9.48_{-9.62}^{+0.38}$\\
020126402 & $\dots$ & $\dots$ & $10.09_{-0.25}^{+0.25}$ & $\dots$\\
020147106 & $8.49$ & $0.04 \pm 0.053$ & $10.10_{-0.13}^{+0.13}$ & $9.41_{-12.27}^{+2.87}$\\
020149061 & $\dots$ & $\dots$ & $10.18_{-0.23}^{+0.23}$ & $11.33_{-14.19}^{+2.86}$\\
020164388 & $8.58$ & $\dots$ & $10.13_{-0.31}^{+0.31}$ & $10.92_{-0.71}^{+0.26}$\\
020167131 & $\dots$ & $\dots$ & $10.08_{-0.20}^{+0.19}$ & $11.58_{-0.61}^{+0.24}$\\
020182331 & $\dots$ & $\dots$ & $10.72_{-0.11}^{+0.11}$ & $11.57_{-0.45}^{+0.22}$\\
020193070 & $8.89$ & $\dots$ & $10.15_{-0.20}^{+0.20}$ & $11.52_{-0.42}^{+0.21}$\\
020208482 & $\dots$ & $\dots$ & $10.17_{-0.16}^{+0.16}$ & $11.92_{-0.43}^{+0.21}$\\
020214655 & $8.76$ & $0.0035 \pm 0.015$ & $10.02_{-0.16}^{+0.16}$ & $10.46_{-1.63}^{+0.30}$\\
020217890 & $\dots$ & $\dots$ & $9.99_{-0.19}^{+0.19}$ & $\dots$\\
020239133 & $\dots$ & $\dots$ & $9.89_{-0.15}^{+0.15}$ & $11.84_{-0.56}^{+0.24}$\\
020240675 & $\dots$ & $\dots$ & $9.96_{-0.18}^{+0.18}$ & $10.26_{-13.12}^{+2.87}$\\
020255799 & $8.80$ & $\dots$ & $9.87_{-0.16}^{+0.16}$ & $8.74_{-11.19}^{+2.46}$\\
020261328 & $8.39$ & $\dots$ & $10.01_{-0.20}^{+0.20}$ & $11.41_{-11.45}^{+0.32}$\\
020278667 & $\dots$ & $\dots$ & $10.28_{-0.16}^{+0.16}$ & $10.95_{-13.71}^{+2.76}$\\
020283083 & $8.74$ & $0.045 \pm 0.026$ & $10.05_{-0.21}^{+0.21}$ & $10.56_{-0.46}^{+0.22}$\\
020283830 & $8.61$ & $\dots$ & $10.37_{-0.17}^{+0.17}$ & $12.02_{-0.30}^{+0.18}$\\
020294045 & $\dots$ & $\dots$ & $9.80_{-0.15}^{+0.15}$ & $12.43_{-0.55}^{+0.23}$\\
020306817 & $\dots$ & $\dots$ & $9.76_{-0.20}^{+0.20}$ & $\dots$\\
020363717 & $8.57$ & $0.0046 \pm 0.039$ & $9.68_{-0.20}^{+0.35}$ & $10.12_{-12.98}^{+2.86}$\\
020370467 & $8.75$ & $-0.022 \pm 0.054$ & $10.57_{-0.14}^{+0.14}$ & $10.35_{-12.70}^{+2.35}$\\
020386743 & $8.53$ & $0.0099 \pm 0.024$ & $9.88_{-0.20}^{+0.20}$ & $10.15_{-0.79}^{+0.26}$\\
020461235 & $8.78$ & $-0.078 \pm 0.023$ & $10.36_{-0.15}^{+0.15}$ & $11.07_{-0.41}^{+0.21}$\\
020461893 & $\dots$ & $\dots$ & $9.66_{-0.21}^{+0.21}$ & $10.60_{-0.54}^{+0.23}$\\
020465775 & $\dots$ & $\dots$ & $10.12_{-0.20}^{+0.20}$ & $10.72_{-0.54}^{+0.23}$\\
140083410 & $8.72$ & $-0.02 \pm 0.041$ & $10.07_{-0.18}^{+0.18}$ & $9.77_{-12.18}^{+2.41}$\\
140096645 & $9.23$ & $\dots$ & $10.40_{-0.23}^{+0.24}$ & $12.75_{-15.00}^{+2.24}$\\
140123568 & $8.75$ & $0.019 \pm 0.074$ & $9.73_{-0.39}^{+0.39}$ & $10.43_{-13.28}^{+2.85}$\\
140137235 & $8.65$ & $-0.051 \pm 0.055$ & $10.07_{-0.29}^{+0.29}$ & $10.67_{-0.36}^{+0.19}$\\
140217425 & $9.02$ & $0.023 \pm 0.0035$ & $10.84_{-0.17}^{+0.17}$ & $12.85_{-0.26}^{+0.16}$\\
140258511 & $8.95$ & $-0.061 \pm 0.037$ & $10.80_{-0.48}^{+0.48}$ & $11.54_{-0.48}^{+0.22}$\\
140262766 & $8.69$ & $0.098 \pm 0.11$ & $9.84_{-0.43}^{+0.43}$ & $11.42_{-14.27}^{+2.86}$\\
140545062 & $8.69$ & $-0.033 \pm 0.012$ & $10.60_{-0.18}^{+0.18}$ & $12.23_{-0.58}^{+0.24}$\\
220014252 & $8.78$ & $0.021 \pm 0.0088$ & $10.78_{-0.21}^{+0.21}$ & $11.57_{-0.50}^{+0.23}$\\
220015726 & $8.75$ & $0.00041 \pm 0.02$ & $10.77_{-0.27}^{+0.27}$ & $12.33_{-14.76}^{+2.43}$\\
220071601 & $\dots$ & $\dots$ & $10.81_{-0.56}^{+0.62}$ & $\dots$\\
220148046 & $\dots$ & $\dots$ & $11.22_{-0.17}^{+0.17}$ & $10.38_{-13.23}^{+2.85}$\\
220376206 & $8.40$ & $0.04 \pm 0.021$ & $10.67_{-0.27}^{+0.27}$ & $12.17_{-0.24}^{+0.15}$\\
220386469 & $8.32$ & $\dots$ & $10.80_{-0.16}^{+0.16}$ & $11.30_{-0.54}^{+0.23}$\\
220397579 & $8.38$ & $-0.022 \pm 0.02$ & $10.23_{-0.17}^{+0.17}$ & $8.18_{-9.16}^{+1.02}$\\
220544103 & $8.89$ & $0.0063 \pm 0.012$ & $10.71_{-0.27}^{+0.27}$ & $11.62_{-0.36}^{+0.20}$\\
220544394 & $8.55$ & $0.034 \pm 0.022$ & $10.34_{-0.23}^{+0.23}$ & $10.54_{-0.44}^{+0.21}$\\
220576226 & $8.75$ & $-0.0073 \pm 0.01$ & $10.31_{-0.23}^{+0.23}$ & $9.77_{-10.09}^{+0.49}$\\
220578040 & $8.64$ & $0.0031 \pm 0.013$ & $10.72_{-0.16}^{+0.17}$ & $12.48_{-14.33}^{+1.86}$\\
220584167 & $8.53$ & $-0.03 \pm 0.013$ & $11.21_{-0.24}^{+0.24}$ & $12.31_{-0.28}^{+0.17}$\\
220596913 & $\dots$ & $\dots$ & $10.68_{-0.30}^{+0.30}$ & $11.69_{-0.11}^{+0.08}$\\
910193711 & $8.77$ & $-0.0091 \pm 0.028$ & $9.99_{-0.18}^{+0.42}$ & $10.56_{-0.42}^{+0.21}$\\
910279515 & $8.60$ & $\dots$ & $10.79_{-0.14}^{+0.14}$ & $12.48_{-0.10}^{+0.08}$\\
\hline
\end{tabular}
\end{table}
\fig{histo_grad} (left panel) shows the distribution of metallicity gradients (in dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${}) for the 26 MASSIV galaxies, for which we were able to measure [N\,\textsc{ii}]6584{} in different regions. The histograms show the full sample (in black), the isolated galaxies (in red) and the interacting ones (in dashed blue), as classified in \S\ref{kin}. For the two latter distributions, we only considered objects classified with a high or medium confidence level (A or B), excluding two isolated galaxies classified with a low confidence level C (VVDS220578040 and VVDS910193711). In the majority of the cases, no clear gradient is detected: the median value of the total sample is $0.0040\pm 0.037$~dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${}. The distributions for isolated (16 objects) and interacting (8 objects) galaxies show similar shapes, with median values $0.0018\pm 0.036$~dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${} for the isolated galaxies, and $0.020 \pm 0.038$~dex kpc$^{-1}$} \newcommand{\car}{$^{-2}${} for interacting ones.
Interestingly though, for twelve of the 26 galaxies, a gradient is detected with a $>1\sigma$ confidence and with nearly the same proportion of positive (seven) and negative (five) gradients (see \fig{grad}). Thus, contrary to the global trend in the local Universe, where the gas-phase metallicity of disk galaxies generally decreases with galactocentric radius, seven of our galaxies have larger metallicities in the outskirts than in the center (where the center is defined as the maximum of the H$\alpha${} flux).
About 1/4 of our sample galaxies thus show a positive metallicity gradient at a $>1\sigma$ confidence level (see blue area in \fig{grad}). Of these, VVDS220376206 displays a positive gradient with $2\sigma$ confidence, while in the galaxies VVDS140217425 and VVDS220014252 a positive gradient of $\sim0.02$ ~dex.kpc$^{-1}$} \newcommand{\car}{$^{-2}${} is detected with $>5\sigma$ confidence. Among the seven galaxies, four are classified as interacting systems, while the three other are isolated. One of the isolated galaxies, the secure case VVDS140217425, appears as a ``chained-galaxy'', with large clumps in the outskirts that could be interpreted as minor mergers. About half (4/8) of the interacting galaxies have a positive metallicity gradient while it concerns only 3/16 ($\sim 20$\%) of the isolated ones. We thus tentatively conclude that the majority of the galaxies showing a positive metallicity gradient are interacting. It is interesting to note also that among the four interacting systems, one galaxy only is classified as a rotating disk.
On the other hand, five galaxies display clear negative metallicity gradients at a $>1\sigma$ confidence level (of which one at $>2\sigma$ and three at $>3\sigma$ confidence level; see \fig{grad} and Table \ref{metal}). Among these five galaxies, three are isolated and two are interacting. Four of these five galaxies are classified as rotating disks.
Although rare, this is not the first time that positive gradients of oxygen abundance were found. Recently, \citet{werk10} reported a positive gradient in a local galaxy and proposed several scenarios to explain their discovery: (i) a radial redistribution of the metal-rich gas produced in the nucleus, (ii) supernovae blowing out metal-rich gas, enriching the IGM, then falling onto the outer parts of the disk, (iii) the result of a past interaction.
At high redshift, \cite{cresci10} have recently studied with SINFONI the metallicity distribution of three Lyman-break galaxies at $z\sim 3$ in the AMAZE/LSD sample. They were able to derive metallicity maps for the three galaxies with high SNR. In each case, they discovered a positive gradient, comparable to the
ones we have found. They favour the scenario in which these positive gradients would be produced by the infall of metal-poor gas into the center of the disks, diluting the gas and lowering its metallicity in the central regions. The authors claim that the discovery of positive gradients in high-redshift disks, pre-selected to be ``isolated'', is a direct evidence for cold gas accretion as a mechanism of mass assembly. Such a conclusion could be balanced arguing that a merger remnant can keep, during a transient phase with a typical timescale of $\sim 0.5$ Gyrs, an inversed metallicity gradient \citep[eg.][]{perez11, torrey11}.
In contrast, our study appears to show that among the seven detected positive gradients, only two galaxies are isolated, the others showing signs of interaction. Cold gas accretion toward the center of disks might thus not be the only process able to lower the central metallicity.
\begin{figure}[t]
\centering
\includegraphics[width=0.5\linewidth]{histo_grad.pdf}\includegraphics[width=0.5\linewidth]{morph.pdf}
\caption{{\it Left}: Metallicity gradients distribution for 26
MASSIV galaxies. The histogram in black line represents the whole
sample, the red and dashed blue ones respectively the isolated and
interacting galaxies distribution. {\it Right}: Metallicity
gradients with respect to the kinematical type (from left to
right, Rotating Isolated, Non-Rotating Isolated, Rotating
Interacting, Non-Rotating Interacting).}
\label{histo_grad}
\end{figure}
\subsection{Relations with global galaxy properties}
\label{sec:behave}
Several authors found correlations between gas-phase metallicity gradients and other global physical parameters of galaxies. The observed metallicity gradients are a function of the morphological type \citep{vilacostas92, marquez02}: they are steep in late-type spirals and almost flat for early-type spirals. Further, in the local Universe the absolute value of the gradients seems to decrease with increasing luminosity (less luminous galaxies have steeper metallicity profiles) as predicted by modeling \citep{prantzos00} and verified by observations \citep{garnett97, vanZee98}.
Our high-redshift sample did not verify the latter correlations and we were not able to test a correlation with the morphological type, as we do not have that information for our MASSIV sample. The right panel of \fig{histo_grad} shows no clear trend between the strength of the metallicity gradient and the kinematical type, nor any correlations with their close environment (isolated or interacting). We can however notice that i) among the isolated objects, the non-rotating galaxies have on average flatter gradients than rotating disks, and ii) the fraction of positive gradients is higher in interacting systems compared with isolated galaxies.
One of the two weak correlations that might be present in our sample is shown in the left panel of
\fig{fig:gradsig}. The strength of the gradient seems to correlate with the velocity dispersion of the galaxy. The latter is derived on beam smearing corrected velocity dispersion maps obtained after velocity field modeling \citep[see][]{epinat11} and thus reflects the true velocity dispersion of the gas.
\begin{figure}[t]
\centering
\includegraphics[width=0.5\linewidth]{grad_vs_sig.pdf}\includegraphics[width=0.5\linewidth]{grad_vs_Z.pdf}
\caption{{\it Left}: Metallicity gradient versus the mean velocity
dispersion of MASSIV galaxies. The black squares are the individual
galaxies, and the red ones represent the median values for 3 bins
of $\sigma_v$, along with the standard deviation in each bin
represented by the error bars. {\it Right}: Metallicity gradients versus
the integrated metallicity of each galaxy, the red line is the best
fit to the data.}\label{fig:gradsig}
\end{figure}
This apparent correlation seems to be driven by the fact that galaxies with high gas velocity dispersion show shallower, often positive, metallicity gradients. This correlation has, to our knowledge, not been observed previously in the local universe, where the velocity dispersion in late-type objects is usually low with $\sigma_v \sim 20$~km s$^{-1}$} \newcommand{\car}{$^{-2}${} \citep{epinat10}. Could turbulent physical conditions in the ISM of high-redshift galaxies be at the origin of the shallow, sometimes positive gradients? This question is difficult to address here, considering the relatively low spatial resolution of our data and the scatter in \fig{fig:gradsig}. At face value, the positive gradients in our $z\sim 1.2$ galaxies might be related to the perturbed physical conditions/motions in the ISM of high-$z$ galaxies, as opposed to the continuous metallicity gradients observed in the relatively quiet ISM of the local spirals.
Finally, we observe as well a tentative anti-correlation between the metallicity gradient of each galaxy and its integrated metallicity (see \fig{fig:gradsig}, right panel). Metallicity gradients are more frequently negative in metal-rich galaxies and more frequently positive in low-metallicity galaxies. If real, this behaviour would support the scenario in which infall of metal-poor gas from the IGM into the center of the disks drives the positive gradients. This infall of pristine gas would be able to reverse the gradient by diluting the central gas metallicity, and lowering the overall metallicity of the galaxy at the same time. The main question would remain: where does the metal-poor gas come from? Accretion of cold gas from the DM reservoir and/or interaction-triggered gas infall/capture from companions?
\subsection{Gas infall rates}
\begin{figure}[t]
\centering
\includegraphics[width=1\linewidth]{mhalo_vs_infallt04.pdf}
\caption{Mean infall rates over 4~Gyr versus the mass of the hosting
DM halo of each MASSIV galaxies. The objects classified as
rotators are circled in black. On the lower left corner is
displayed a typical errorbar on the DM halo mass.}
\label{fig:infall}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=1\linewidth]{infall_vs_sfr_erreur.pdf}
\caption{Mean infall rates versus the star formation rates derived
with the H$\alpha${} luminosity and corrected for extinction. The line is the best fit to
the data.}
\label{fig:infall_sfr}
\end{figure}
\label{sec:toy}
As suggested already by previous studies \citep[eg.][]{werk10, rupke10}, during an interaction, metal-poor gas from the outskirts of the galaxy could
radially flow towards the center and dilute the metallicity in the inner, high star-forming regions. At the same time, metal-enriched gas
could be transported to the outer parts during the interaction, overall resulting in a flattening of the metallicity gradient.
The two scenarios explaining the positive metallicity gradients \--- cold gas accretion and gas redistribution during interactions \--- have
in common that metal-poor gas needs to be transported efficiently to the center of the objects on time scales shorter than the star formation.
We consider the possibility of infall of pristine gas onto the disk in the context of a chemical evolution model in order to explain the
positive gradients. Our toy model for the chemical evolution of galaxies assumes (i) an instantaneous recycling approximation (IRA), (ii) the
infall of metal-free gas onto the disks, and (iii) radial flows of gas into the disks. This leads to the following equations:
\begin{eqnarray}
\frac{\partial g}{\partial t} + \frac{\partial }{r \partial r} (r v g) & = & -(1-R)\psi(t) + f\label{toy1}\\
\frac{\partial Z}{\partial t} + \frac{\partial }{r \partial r} (r v Z ) & = & y(1-R)\psi(t) - z(1-R)\psi(t)\label{toy2}
\end{eqnarray}
where $g$ is the gas surface density, $\psi$ is the star formation rate per unit area, $R$ is the ``returned fraction'', $f$ is the infall rate
(per unit area) of intergalactic metal-free gas, $Z = z g$ is the metal content, $y$ is the stellar yield, and $v$ is the velocity of the radial
gas flow (positive toward the outer parts). Following a Schmidt-Kennicutt law for the star formation rate density
($\Sigma_\psi = C g^n$, $C = 1.6\times 10^{-27}$, $n=0.7$) we solve these equations for each galaxies in our sample, assuming typical
values for $C$, $n$, $y$, $v$ and $R$ (see above and below). After combining equations (\ref{toy1}) and (\ref{toy2}), we end up with the
following equation allowing to derive the infall parameter $f$:
\begin{equation}
\label{eqz}
\frac{\partial z}{\partial t} + v\frac{\partial z}{\partial r} = y(1-R) C^{1/n} \psi^{1-1/n} - z C^{1/n} \psi^{-1/n} f
\end{equation}
Assuming the physical parameters involved in eq.~(\ref{eqz}) to remain constant for over $\sim 4$~Gyr (i.e.~valid for galaxies at $z \simeq 1.2$),
we derive a mean infall rate per unit area that can be integrated since the formation epoch. To this end, we use a true yield value of $y=0.019$,
and a returned fraction of $\sim 40$\%: $R = 0.4$ \citep{erb08}. The radial flow velocity of high-redshift galaxies is known to be higher than in local
spirals (e.g.~$\sim 1$~km s$^{-1}$} \newcommand{\car}{$^{-2}${} in the case of the Milky Way). This velocity can be approximated to be the velocity towards the center of the migrating
clumps \--- in which the star formation takes place \--- which then leads to velocities of $10-50$~km s$^{-1}$} \newcommand{\car}{$^{-2}${} \citep{bournaud07}. For our purpose,
we used for each MASSIV galaxy the estimated velocity dispersion reported in \cite{epinat11}.
For the star formation rates, we used those derived from the H$\alpha${} luminosity, corrected for extinction, and listed in Table~\ref{eml}.
Although the H$\alpha${} luminosity only reflects the instantaneous star formation rate, we assume it to be an average value
since the formation epoch. As our galaxies were selected to be star-forming, this is most likely an upper limit on the average star formation.
The so-calculated global infall rates range from a few to several hundreds $\mathrm{M_\odot yr^{-1}}$. These values are high compared to the
output of hydro-dynamical simulations of typical disk galaxies (50~M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}$ yr$^{-1}$} \newcommand{\car}{$^{-2}${}). However, the rate of infalling gas is believed to depend on the size and
mass of the host DM halo, being able to reach a few hundreds of M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}$ yr$^{-1}$} \newcommand{\car}{$^{-2}${} for the most massive galaxies \citep{keres05, erb08}.
We have therefore derived DM halo masses for each galaxy in our sample (see Section \ref{sec:masses}) and compared it to our calculated infall
rates (see \fig{fig:infall}). As expected there is a tendency for galaxies to accrete more gas in the most massive halos. We further divided the sample in
three bins with respect to the (extinction-corrected) star formation rate. In that case, we noticed that galaxies with the highest SFR also have the highest
gas-infall rate, whereas galaxies with the lowest SFR show infall rates typical for local galaxies ($\sim$50~M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}$ yr$^{-1}$} \newcommand{\car}{$^{-2}${}).
Two MASSIV galaxies deviate from the global trend in \fig{fig:infall}. VVDS220397579, located at the left end of the graph, has a very low DM
halo mass but an important infall rate ($\sim 650~\mathrm{M_\odot yr^{-1}}$), while VVDS220376206 has by far the highest infall ($>10^3~\mathrm{M_\odot yr^{-1}}$)
and star formation ($>250~\mathrm{M_\odot yr^{-1}}$) rates . For VVDS220397579 it is very likely that its DM halo mass has been underestimated. Indeed, this
galaxy could be a face-on disk as the kinematical modeling returned a very low rotation velocity of $\sim 9$~km s$^{-1}$} \newcommand{\car}{$^{-2}${}. As the DM halo mass is directly proportional
to the maximum rotation velocity (see equation \ref{eq:mhalo}) the DM halo mass for VVDS220397579 is most probably only a lower limit.
In \fig{fig:infall}, we have circled the objects which, according to our kinematical classification, are isolated and rotating. For those objects the DM halo
mass is more reliable, and they appear indeed to have the largest DM halo masses. In order to investigate the relationship between the infall rate and the
star formation rate, we show in \fig{fig:infall_sfr} the mean infall rates against the SFRs corrected for extinction. As expected from equation~\ref{eqz}, a clear linear
correlation between these two quantities can be seen: the infall rates appear to be directly proportional to the star formation rates. The best fit to the data gives
a slope of $2.71$ for this relation, which is defined as the infall parameter $f_i$. This was already shown to be a typical value for galaxies at these redshifts \citep{erb08}.
\section{Summary and conclusions}
\label{sec:conclu}
This paper presents the chemical abundance analysis of the ``first epoch'' sample (50 galaxies) of the MASSIV survey. Complementary analyses focused on the kinematical classification and the fundamental scaling relations can be found in two companion papers \citep{epinat11, vergani11}, as well as a description of the whole MASSIV sample \citep{contini11}.
We have been able to measure emission lines ([N\,\textsc{ii}]6584{} and H$\alpha${}) in 34 integrated spectra of MASSIV galaxies. Within this sample, we have identified one galaxy hosting an AGN. Chemical abundance estimates could be obtained for the remaining 33 star-forming galaxies.
For 26 galaxies, we have been able to derive a metallicity gradient, defining annular regions around the peak of the H$\alpha${} flux (which corresponds, in most cases, to the kinematical center). While just over half of our sample is compatible with a zero metallicity gradient, about a quarter of the galaxies shows {\it positive} gradients: the metallicity increases from the center to the outer parts of the galaxy. Among these latter (seven) galaxies, four are classified as interacting systems, one is probably a chain galaxy, and two are classified as isolated.
Flat, more rarely positive, metallicity gradients have already been found in interacting galaxies in the local Universe. They are explained by the infall of metal-poor gas onto the central parts during the encounter \citep{werk10, rupke10}.
Two of our galaxies showing a positive metallicity gradient are classified as isolated and do not show any sign of recent interaction. Three analogue objects were reported as isolated disk galaxies at $z\sim3$ by \citet{cresci10}. In these cases, cold gas accretion onto the central regions of the disks seems to be the most plausible scenario. Even if tentative, there is a very interesting trend as a function of redshift. At $z\sim 3$ almost all isolated galaxies have a positive gradient, whereas this fraction drops to $\sim 15-20$\% at $\sim1.2$ and is almost equal to zero in the local universe. If cold accretion is the main process to explain the positive metallicity gradients in isolated disks, that suggests that the epoch where cold accretion dominates the mass assembly processes is at $z\geq 2$.
We noticed in our sample a tendency for galaxies with the highest gaseous velocity dispersion to have a shallow/positive gradient, which highlights the different physical conditions observed in the ISM of high-redshift galaxies (high velocity dispersion compared to local spirals). We discovered a weak correlation between gradient and global metallicity of the galaxy: metal-poor galaxies preferentially have flat/positive gradients while metal-rich ones tend to display negative metallicity gradients. This behaviour can also be explained by the infall of
metal-poor gas onto the center of the disks, diluting the overall metallicity.
Finally, applying a simple chemical evolution model with radial flows of gas, we estimated infall rates of pristine gas onto the disks. We found values up to several hundred of M$_{\odot}$} \newcommand{\arsec}{$^{\prime\prime}${} per years, and a tendency for the maximum infall rate to increases with the DM halo mass.
The analysis of the spatially-resolved metallicity of galaxies will be further extended to the full MASSIV sample, enabling a better statistics and hence a stronger interpretation in terms of galaxy assembly scenarios. The final sample will
also allow to establish the mass-metallicity and fundamental metallicity relations at $0.9 < z < 1.8$, adding constraints on the evolutionary status of star-forming galaxies at high redshifts.
\begin{acknowledgements}
We wish to thank L\'eo Michel-Dansac and Fr\'ed\'eric Bournaud for
their help and useful comments on this work. We thank also the referee for useful suggestions. This work has been partially supported by the
CNRS-INSU and its Programme National Cosmologie-Galaxies (France) and by the french ANR grant ANR-07-JCJC-0009. DV acknowledges the support from the INAF contract PRIN-2008/1.06.11.02.
\end{acknowledgements}
\newpage \bibliographystyle{aa} |
\section{Introduction}
From the observation of neutrino oscillations, we now know that lepton
flavors are not conserved. However,
the mixing and small neutrino mass differences seen in oscillations
have a negligible effect on charged-lepton flavor violating (CLFV) reactions.
Thus, the CLFV
reactions provide a discovery window for interactions beyond Standard
Model expectations \cite{Kuno:1999jp,LeptMomBook}.
Muons play a central role in searches for CLFV
\cite{Kuno:1999jp,LeptMomBook}, because they can be produced in large numbers
and live relatively long.
One reaction that can be probed with particularly high sensitivity
is coherent muon-electron conversion in a muonic atom,
\begin{equation}
\mu^{-}+(A,Z)\to e^{-}+(A,Z),\label{eq:mueconv}
\end{equation}
where $(A,Z)$ represents a nucleus of atomic number $Z$ and mass
number $A$. It has the advantage of producing just a single particle, a mono-energetic electron. It does not have the problem of accidental background that plagues searches for the decay $\mu^+\to e^+\gamma$, which can be mimicked by a positron from a normal muon decay and a photon
coming from the radiative decay of a different muon, bremsstrahlung,
or positron annihilation-in-flight.
Various experiments have been performed over the years
to search for the conversion \cite{Marciano:2008zz}. The most stringent results
come from the SINDRUM II Collaboration \cite{Bertl:2006up},
which reports an upper limit of $7\times10^{-13}$ for the branching
ratio of the conversion process relative to muon capture in gold. Several
new efforts are being planned. In the nearest future, the DeeMe
Collaboration \cite{DeeMe} has proposed to reach $10^{-14}$ sensitivity. Larger scale searches,
Mu2e at Fermilab \cite{Carey:2008zz} and COMET at J-PARC \cite{Cui:2009zz},
aim for sensitivities below $10^{-16}$. In the long run, intensity
upgrades at Fermilab and the proposal PRISM/PRIME at
J-PARC may allow them to reach $10^{-18}$ sensitivity. A quite remarkable improvement
of about four orders of magnitude, with respect to the current limit, is
therefore envisaged.
The success of the conversion searches depends critically on control
of the background events. The signal for the $\mu-e$ conversion process
in Eq.~(\ref{eq:mueconv}) is a mono-energetic electron with energy
$E_{\mu e}$, given by
\begin{equation}
E_{\mu e}=m_{\mu}-E_{\mathrm{b}}-E_{\mathrm{rec}},\label{eq:conven}
\end{equation}
where $m_{\mu}$ is the muon mass, $E_{\mathrm{b}}\simeq Z^2\alpha^2m_{\mu}/2$ is the binding
energy of the muonic atom, and $E_{\mathrm{rec}}\simeq m_{\mu}^2/(2m_N)$ is the nuclear-recoil
energy, with $\alpha$ the fine-structure constant and $m_N$ the
nucleus mass. The main physics background for this signal comes from the so-called
muon decay in orbit (DIO), a process in which the muon decays in
the normal way, $\mu^{-}\to e^{-}\overline{\nu}_{e}\nu_{\mu}$, while
in the orbit of the atom. Whereas in a free muon decay, in order to
conserve energy and three-momentum, the maximum electron energy is
$m_{\mu}/2$, for DIO, the nucleus recoil can balance the electron's
three-momentum taking basically no energy. This allows for the
maximum electron energy to be $E_{\mu e}$, close to the full muon mass $m_\mu$.
Therefore, the high-energy electrons from the muon decay in orbit constitute a
background for conversion searches.
\section{Muon decay in orbit}
Several theoretical studies of the muon decay in orbit have been
published. Expressions describing the electron spectrum including relativistic effects in the muon
wavefunction, the Coulomb interaction between the electron and the
nucleus and a finite nuclear size have been available for some time
\cite{Haenggi:1974hp,Watanabe:1987su,Watanabe:1993}. However, the high-energy
endpoint of the spectrum, which is the most important region for conversion-search experiments, was not studied in detail.
References \cite{Shanker:1981mi,Shanker:1996rz} did study the high-energy
end of the electron spectrum, and presented approximate results which
allow for a quick rough estimate of the muon decay in orbit contribution
to the background in conversion experiments. We have performed a new
evaluation of the DIO spectrum, considering in detail all the
effects needed in the high-energy region \cite{Czarnecki:2011mx}.
Our results describe the background contribution for $\mu-e$
conversion searches, as well as a check on previous low- and
high-energy partial calculations \cite{Watanabe:1993,Shanker:1981mi}
and an interpolation between them. It is worth emphasizing that not
only the high-energy region is relevant for conversion experiments, but
the full spectrum is necessary in order to study reconstruction errors
in the detector.
To obtain the correct result for the high-energy tail of the spectrum
it is crucial to include nuclear-recoil effects, since they modify the
endpoint energy (see Eq.~(\ref{eq:conven})). Also, to produce an
on-shell electron with energy around $m_{\mu}$, either the muon must be at the
tail of the bound-state wavefunction or the produced electron must interact
with the nucleus. This tells us that the full Dirac
equation for the muon as well as the interaction of
the outgoing electron with the field of the nucleus must be taken into
account. Also, finite-nuclear-size effects will be most
important in this region. Order $\alpha$ radiative
corrections are not expected to significantly modify the results at
the endpoint, and are not included in our results. Uncertainties in
the modelling of finite nuclear-size effects induce errors in the
spectrum that increase as we approach the endpoint, but those errors are never
larger than a few percent.
\section{Results and discussion}
Here we present our results for the elements that are relevant for the
upcoming conversion experiments. The DeeMe Collaboration plans to use a
silicon-carbide target, whereas the Mu2e and COMET Collaborations are
considering aluminum and titanium as targets. In Table \ref{tab:en} we give the values of
the bound muon energy $E_{\mu}=m_{\mu}-E_{\mathrm{b}}$ and the electron
endpoint energy $E_{\mu e}$, for carbon, aluminum, silicon and titanium.
\begin{table}
\caption{Values for muon energies $E_{\mu}$, nuclear masses $m_N$, and
endpoint energies $E_{\mu e}$.}
\label{tab:en}
\begin{tabular}{cllll}
\hline\noalign{\smallskip}
Nucleus & Z & $E_{\mu}$ (MeV) & $m_N$ (MeV )& $E_{\mu e}$ (MeV) \\
\noalign{\smallskip}\hline\noalign{\smallskip}
C & 6 & 105.557 & 11188 & 105.06 \\
Al &13 & 105.194 & 25133 & 104.973 \\
Si & 14 & 105.121 & 26162 & 104.91 \\
Ti & 22 & 104.394 & 44588 & 104.272 \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
We present the results of the numerical evaluation of the spectra for
those elements in Figs.~\ref{fig:eplog} and \ref{fig:full}.
\begin{figure}
\includegraphics[width=\textwidth]{endpoint.eps}
\caption{Electron spectrum, normalized to the free-muon decay rate
$\Gamma_0$. The solid blue line is for carbon, the black dotted line
for aluminum, the green dot-dashed line for silicon and the red dashed
line for titanium.}
\label{fig:eplog}
\end{figure}
\begin{figure}
\includegraphics[width=0.49\textwidth]{log.eps}
\includegraphics[width=0.49\textwidth]{linear.eps}
\caption{Electron spectrum for the full range of $E_e$; see caption of Fig.~\ref{fig:eplog}.}
\label{fig:full}
\end{figure}
Regarding the nuclear distributions, a two-parameter Fermi
distribution has been used for aluminum and titanium, a
three-parameter Fermi distribution for silicon and a Fourier-Bessel
expansion for carbon \cite{Czarnecki:2011mx,De Jager:1987qc,Fricke:1995zz}.
Those results are useful for assessing the DIO background events from carbon for the
DeeMe experiment which will search primarily for conversion in silicon.
They also illustrate the electron resolution requirements, as a function of stopping target,
needed to reach future high sensitivity goals.
The high sensitivity that the upcoming conversion experiments will
reach may also allow them to improve the present bounds on some exotic muon
decays, like the decay of the muon into an electron and a majoron (a
Goldstone boson that appears in models where lepton number is a
spontaneously broken global symmetry) \cite{Tormo:2011et}.
\begin{acknowledgements}
This research was supported by Science and
Engineering Research Canada (NSERC) and by the United States Department of
Energy under Grant Contract DE-AC02-98CH10886.
\end{acknowledgements}
|
\section{Introduction}
The Tunable Diode Laser Absorption Spectroscopy (TDLAS) is a sensitive and fast method for gas sensing used for instance to measure concentrations of polluting gas such as CH$_4$ and CO.
This technique requires the use of a tunable, stable and single mode operating source emitting in the mid infrared wavelength range beyond \unit{2}{\micro\meter} where strong absorption lines are present for these gas~\cite{chen_apb_2010, boehm_jcg_2010}. VCSELs are highly suitable for this kind of application, however, the development of continuous wave devices operating at room temperature at this wavelength range is still an important challenge.
Due to the small thickness of the gain region, VCSEL structures require high quality cavities with highly reflective Bragg mirrors (R~$>$~99 \%). In the mid-infrared wavelength range, even though the best performances were obtained with AlGaInAsSb material system, the relatively low index contrast ($\Delta n \sim$~0.5) between DBR layers make the mirrors become as thick as \unit{11}{\micro\meter} impairing the electro-thermal-optical properties of the structure~\cite{cerutti_jcg_2009}. Currently, laser emission has been shown from an all-epitaxial monolithic microcavity near \unit{2.3}{\micro\meter} in a continuous wave mode and in quasi-CW (5~\%, \unit{1}{\micro\second}) up to \unit{2.63}{\micro\meter}~\cite{ducanchez_el_2009_2}. Another hybrid structure made of a dielectric top mirror and a buried tunnel junction operates in CW emission at room temperature in the 2.4-\unit{2.6}{\micro\meter} wavelength range~\cite{bachmann_njp_2009}.
One solution to increase the emission wavelength is the use of a high contrast grating mirror as top cavity mirror~\cite{huang_np_2007}. Such gratings, combined with a low index sub-layer, can exhibit high reflectivity of more than 99.9~\% for bandwidths larger than 100~nm~\cite{mateus_ptl_2004}. Moreover, due to their one dimensional symmetry, a high polarization selectivity can be performed by these mirrors and with a total thickness of less than \unit{2}{\micro\meter}~\cite{chevallier_apa_2010}, the stability and quality of the VCSEL emitted beam should be improved.
However, contrary to Bragg mirrors, the explanation of the optical response of high contrast grating (HCG) is not straightforward and the design adjustment made in order to achieve the required properties for a VCSEL application is still complex. Even though several formalism have been defined, based for instance on the destructive interference of modes~\cite{karagodsky_oe_2010}, the use of an optimization algorithm for the design combined with a numerical computation of the reflectivity keeps the advantage of versatility. Indeed, such a method easily allows the user to aim for specific properties of the HCG. It has been used for instance to design wide band and high diffraction efficiency grating~\cite{wang_ol_2010}, large bandwidth grating mirror for both TE and TM polarization~\cite{shokooh-saremi_oe_2008} or, like in this work, polarization selective large bandwidth mirror with technological constraints on the design dimensions~\cite{chevallier_apa_2010}.
In this work, we present a Si/SiO$_2$ HCG design optimized for a VCSEL application at $\lambda = $~\unit{2.65}{\micro\meter}. Then, through a precise study of the computed tolerances of the structure dimensions, the robustness is enhanced with respect to the fabrication errors.
\section{Design of the HCG structure}
The high contrast grating structure studied in this work is based on Si/SiO$_2$ materials which properties and fabrication process are well known. The grating is made of silicium (n~=~3.435) on top of a low index layer of SiO$_2$ (n~=~1.509) used to achieved high reflectivity~\cite{mateus_ptl_2004} and allowing the use of a selective etching method for the fabrication of the grating. In order to increase the total reflectivity of the reflector and broaden the stopband, two quarter wavelength layers of Si/SiO$_2$ are combined with the grating~\cite{chevallier_apa_2010}. For the simulation of the design, the substrate is chosen as the VCSEL cavity material with an optical index of n~=~3.521.
In order to be well suitable for a VCSEL application, HCGs have to exhibit optical properties as we have defined with a 99.9~\% transverse magnetic (TM) reflectivity for the largest possible bandwidth. Moreover, to ensure a polarization stability of the emitted beam, the reflectivity of the transverse electric (TE) mode has been chosen to be kept lower than 90 \% for the whole bandwidth.
A 99.5~\% TM reflectivity should be enough to achieve laser emission but a 0.4~\% security margin has been chosen to account for possible experimental growth imperfections and losses due to absorption. Indeed, the presence of OH radicals as impurities in SiO$_2$ results in an absorption band in the 2.6-\unit{2.9}{\micro\meter} range~\cite{soref_joa_2006}. In this reference, the absorption of silica has been measured with a 10 dB/cm value at \unit{2.65}{\micro\meter}, i.e. the refractive index becomes $n_{SiO2} = 1.509 + 2.1e^{-4} j$, which results in a 0.1~\% fall of the HCG reflectivity while a 0.4~\% decrease of reflectivity has been observed for a simulation with a 20 dB/cm attenuation coefficient ( $n_{SiO2} = 1.509 + 1.7e^{-3} j$ ) at \unit{2.675}{\micro\meter}. This latter value appears therefore as the maximum absorption value allowed for the studied structure. However, the OH impurities concentration should strongly depends on the fabrication process and the exact value of the absorption must be determined in each case. For this theoretical study, a pure SiO$_2$ without any OH absorption was considered. Nevertheless, if the silica absorption is proven to be too high, this material can be replaced by another dielectric such as Si$_3$N$_4$.
The evaluation of the reflectivity of the mirror is made thanks to a rigorous coupled wave analysis (RCWA)~\cite{mrcwa, moharam_josaa_1995} which numerically finds an exact solution of Maxwell's equations for the electromagnetic diffraction problem of an infinite grating structure.
Reflection spectra for TE and TM modes are thus computed and through a well defined quality factor~\cite{chevallier_apa_2010}, the optical performance of the reflector is numerically evaluated. This performance is then maximized by a genetic optimization algorithm which searches for a global maximum of the quality factor by adjusting the design. The use of a global algorithm~\cite{openopt} is mandatory in this problem since the quality factor function exhibits numerous local maxima.
\begin{figure}
\center\includegraphics[width=4.5cm]{Fig1}
\caption{Scheme of the structure. The silicium grating defined by the empty length of the grooves $L_e$, the filled length $L_f$ and the grating thickness $T_g$ is on top of a SiO$_2$ layer of thickness $T_L$. These four parameters are adjusted to meet the characteristics of a VCSEL reflector.}
\label{scheme}
\end{figure}
In this work, the performances of the grating are optimized by adjusting the empty and filled lengths ($L_e$ and $L_f$), the grating thickness $T_g$ and the sublayer thickness $T_L$ as shown on Figure \ref{scheme}.
The use of an automated optimization presents the advantage of easily imposing technological constraints on the dimensions of the structure. These constraints have been set on the lengths $L_e$ and $L_f$ limited to a minimum value of 500~nm which should ease the photolithographic process. Besides, the etching process is also a critical step of the grating fabrication and, to achieve the theoretical grating profile presented on Figure~\ref{scheme}, the shape factor $SF = L_e / T_g$ of the grooves is kept at a minimum value of 0.9 since squared patterns are easier to etch than deep ones.
\begin{figure*}
\center\includegraphics[width=10cm]{Fig2}
\caption{Reflection spectra for the TM mode (blue) and TE mode (dashed red) of the structure automatically optimized by a genetic-based algorithm. This optimum design exhibits a 307~nm large bandwidth with a 99.9~\% high TM reflectivity.}
\label{optimum}
\end{figure*}
The optimization for $\lambda = $~\unit{2.65}{\micro\meter} of the structure shown on Figure \ref{scheme} results in a 307~nm large bandwidth mirror exhibiting a high polarization selectivity with $R_{TE} < $~90 \% (Fig.~\ref{optimum}). The optimum lengths of the design are given in Table~\ref{table1} and satisfy the technological constraints with $L_e$~=~799~nm, $L_f$~=~541~nm and a shape factor $SF$~=~0.9. The grating thickness $T_g$~=~886~nm and the sublayer thickness $T_L$~=~321~nm combined with the two quarter wavelength layers result in a \unit{1.84}{\micro\meter} thick reflector.
Even though the performances of this HCG are well adapted for a VCSEL application, the tolerances of the dimensions of this grating are critical. For instance, the 884~nm minimum grating thickness leads to a maximum error allowed as small as 2~nm.
\section{Optimization of the robustness}
\begin{table*}
\caption{Tolerances of the optimum design found by the genetic-based algorithm.}
\label{table1}
\begin{center}
\setlength{\tabcolsep}{0.5cm}
\begin{tabular}{llll}
\toprule\noalign{\smallskip}
& Optimum & Minimum & Maximum\\
\noalign{\smallskip}\hline\noalign{\smallskip}
$L_e$ & 799 nm & 689 nm & 803 nm \\
$L_f$ & 541 nm & 495 nm & 629 nm \\
$T_g$ & 886 nm & 884 nm & 932 nm \\
$T_L$ & 321 nm & 196 nm & 414 nm \\[0.2cm]
$\Lambda = L_e + L_f$ & 1340 nm & 1260 nm & 1345 nm \\
$FF = L_f / \Lambda$ & 40.37 \% & 35.97 \% & 45.67 \%\\[0.2cm]
$\alpha_e = L_e + T_g$ & 1685 nm & 1652 nm & 1689 nm \\
$\beta_e = L_e / \alpha_e$ & 52.58 \% & 50.80 \% & 54.89 \%\\[0.2cm]
$\alpha_f = L_f + T_g$ & 1427 nm & 1388 nm & 1430 nm \\
$\beta_f = L_f / \alpha_f$ & 62.09 \% & 59.13 \% & 64.74 \%\\
\noalign{\smallskip}\bottomrule
\end{tabular}
\end{center}
\end{table*}
The evaluation of the tolerances is performed by computing the variation range of one parameter for which the HCG keeps a 99.9~\% TM reflectivity together with a $R_{TE} <$~90~\% at $\lambda_0 =$~\unit{2.65}{\micro\meter}. This computation is made by varying one parameter at a time, for instance the grating thickness $T_g$, while keeping the other ones ($T_L$, $L_e$ and $L_f$) constant at their optimal value.
The computation of the tolerances of the design found by the optimization algorithm exhibits large variation ranges of $\Delta L_e =$~14~\%, $\Delta L_f =$~25~\%, $\Delta T_g =$~5~\% and $\Delta T_L =$~68~\%. However, the optimum point is not centred in these variation ranges and the real tolerance values are much more limited with for instance a maximum increase of 4~nm on the empty length $L_e$. Moreover, these variation ranges are computed separately and give no information of the simultaneous error allowed on several parameters.
A solution to access this information is to compute the variation range of combinations of two dimensions. In the following, only ($L_e$,~$L_f$), ($L_e$,~$T_g$) and ($L_f$,~$T_g$) combinations are evaluated since the sublayer thickness $T_L$ is the most tolerant and centred parameter. Besides, $T_L$ does not depend on the etching process and should be fabricated with a better accuracy than the grating parameters.
\begin{figure}
\center\includegraphics[width=8cm]{Fig3}
\caption{Tolerance map of $L_e$ and $L_f$. The variation ranges of the grating period $\Lambda = L_e + L_f$ and the fill factor $FF = L_f/\Lambda$ define a polygon (grey) of allowed ($L_e,~L_f$) couples for the design. The centre of the incircle ($\star$) enhances the robustness of the optimum point ($\circ$).}
\label{LeLf}
\end{figure}
For the first couple ($L_e$,~$L_f$), the variation ranges of two significant combinations of $L_e$ and $L_f$ are evaluated. The first one is the grating period given by $\Lambda = L_e + L_f$. This parameter is varied around its optimum value of $\Lambda = $~1340~nm while keeping the fill factor defined as $FF = L_e / \Lambda$ at its optimum value of 40.37~\%. The second combination of $L_e$ and $L_f$ is the fill factor $FF$ which variation range is evaluated with a constant grating period of $\Lambda = $1340~nm. Large tolerance values are exhibited by these parameters with 85~nm for $\Delta \Lambda$ and 9.7~\% for $\Delta FF$. The minimum and maximum of the variation ranges of the four parameters $L_e$, $L_f$, $\Lambda$ and $FF$, summarized in Table \ref{table1}, creating eight ($L_e$,~$L_f$) couples which can be plotted in a $L_f$ versus $L_e$ graph (Fig.~\ref{LeLf}). In the ($L_e$,~$L_f$) plan, these points define a polygon, in grey on Figure~\ref{LeLf}, which area represents the tolerance area of ($L_e$,~$L_f$)~couples allowed for the design.
The representation of the tolerances in a ($L_e$,~$L_f$) plan also shows very well the position of the optimum point within the tolerance area. In this case, the optimum point ($\circ$ on Figure~\ref{LeLf}) is located at an edge of the tolerance area ($L_e$~=~799~nm, $L_f$~=~541~nm). In order to increase the robustness regarding to the errors of fabrication which could be made on $L_e$ and $L_f$, the design should be centred within the tolerance area. To do so, the centre of the largest incircle of the polygon representing the tolerance area is computed. This point is thus the farthest point from any edge of the polygon. The centre ($\star$) on Figure~\ref{LeLf} corresponds to $L_e =$~753~nm and $L_f =$~560~nm and with a incircle radius of 47~nm, ensures a minimum tolerance of 94~nm on any combination of ($L_e$,~$L_f$).
\begin{figure}
\center\includegraphics[width=8cm]{Fig4}
\caption{Tolerance map of $L_e$ and $T_g$. The variation ranges of $\alpha_e = L_e + T_g$ and $\beta_e = L_e/\alpha_e$ define a polygon (grey) of allowed ($L_e,~T_g$) couples for the design. The centre of the incircle ($\star$) enhances the robustness of the optimum point ($\circ$).}
\label{LeTg}
\end{figure}
The second dimension couple studied is the empty length and grating thickness ($L_e$,~$T_g$). In this case, both values of the optimum point (799,~886) are closed to the tolerance limit ($L_{e~max}$~=~803~nm, $T_{g~min}$~=~884~nm). This can easily be seen on Figure~\ref{LeTg} where the optimum is localized at the edge of the tolerance area. In this map, the first combination of ($L_e$,~$T_g$) is defined by $\alpha_e = L_e + T_g$ and the second one by $\beta_e = L_e/\alpha_e$. The polygon defined by the height extrema of the variation ranges of $L_e$, $T_g$, $\alpha_e$ and $\beta_e$ (Table~\ref{table1}) exhibits an incircle with a radius of 28~nm centred at (772, 898) making a tolerance of a minimum value of 56~nm on the lengths ($L_e$,~$T_g$) and any kind of their combinations.
The last couple, formed by the filled length $L_f$ and grating thickness $T_g$, is optimized by computing the tolerance values of $\alpha_f = L_f + T_g$ and $\beta_f = L_f / \alpha_f$. As shown on Figure~\ref{LfTg}, the polygon exhibits two large areas joined by a very thin path where the optimum point is located at (541,~886). Once again, the optimum point is in a critical location closed to several edges of the tolerance area. By choosing the centre of the incircle of the polygon, located at (519,~894), variation ranges of the lengths $L_f$ and $T_g$ can be increased to a 44~nm value.
\begin{figure}
\center\includegraphics[width=8cm]{Fig5}
\caption{Tolerance map of $L_f$ and $T_g$. The variation ranges of $\alpha_f = L_f + T_g$ and $\beta_f = L_f/\alpha_f$ define a polygon (grey) of allowed ($L_f,~T_g$) couples for the design. The centre of the incircle ($\star$) enhances the robustness of the optimum point ($\circ$).}
\label{LfTg}
\end{figure}
\section{Characteristics of the robust HCG}
\begin{table}
\caption{Tolerances of the resulting design with optimized variation ranges.}
\label{table2}
\begin{center}\setlength{\tabcolsep}{0.3cm}
\begin{tabular}{lllll}
\toprule\noalign{\smallskip}
& Optimum & Minimum & Maximum & \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$L_e$ & 773 nm & 703 nm & 821 nm & (15 \%)\\
$L_f$ & 541 nm & 504 nm & 608 nm & (19 \%)\\
$T_g$ & 894 nm & 870 nm & 924 nm & (6 \%)\\
$T_L$ & 321 nm & 0 nm & 674 nm & (210 \%)\\
\noalign{\smallskip}\bottomrule
\end{tabular}
\end{center}
\end{table}
\begin{figure*}
\center\includegraphics[width=10cm]{Fig6}
\caption{Reflection spectra for the TM mode (blue) and TE mode (dashed red) of the robust design with optimized tolerance values. This mirror exhibits a 250~nm large bandwidth for $R_{TM} >$~99.9~\% and between 5~\% and 210~\% of tolerance on its dimensions.}
\label{centre}
\end{figure*}
As a result of this optimization of the tolerances, three different designs are obtained. Each one is optimized to have the best tolerances for the couples ($L_e$,~$L_f$), ($L_e$,~$T_g$) and ($L_f$,~$T_g$). These three points represent a triangle in a three dimensional space ($L_e$,~$L_f$,~$T_g$). Thus, to find a unique design with optimized tolerances, the centre of the incircle of this triangle has been computed and result in a final optimized design with lengths $L_e = $~773~nm, $L_f = $~541~nm and $T_g = $~894~nm. This structure exhibits a 250~nm large bandwidth as shown on Figure~\ref{centre}, which is 59~nm less than the optimum one found by the genetic-based algorithm. However, the computation of the variation ranges of the design dimensions (Table~\ref{table2}) exhibits a very robust design with $\Delta L_e = $~15\%, $\Delta L_f = $~19\%, $\Delta T_g = $~6\% and $\Delta T_L = $~210\%. Each optimum dimension of the structure is well centred within the variation ranges which would ease the fabrication process.
\section{Conclusion}
In this work, a Si/SiO$_2$ high contrast grating mirror has been designed for a VCSEL application at~\unit{2.65}{\micro\meter}. This mirror exhibits a 250~nm large bandwidth for $R_{TM} >$~99.9~\% together with a strong polarization selectivity by keeping $R_{TE} <$~90~\%. Moreover, with a total thickness of less than \unit{2}{\micro\meter}, such reflector should improve the quality of the emitted laser beam of the VCSEL.
From a technological point of view, the fabrication constraints are respected with a large pattern resolution ($>$~500~nm) and a grating shape factor closed to 0.9. Moreover, the robustness with respect to the fabrication errors has been enhanced and leads to tolerance values larger than 5~\% on the structure dimensions. Such characteristics make this HCG well adapted for a VCSEL application and should limit the pitfalls during the manufacturing process. Moreover, as spotted in~\cite{mateus_ptl_2004}, HCG can be scaled with wavelength in the limit of the refractive index dispersion.
\section*{Acknowledgements}
The authors thank the French ANR for financial support in the framework of Marsupilami project (ANR-09-BLAN-0166-03) and IES and LAAS (France), partners of LMOPS/Sup\'elec in this project. This work was also partly funded by the InterCell grant (http://intercell.metz.supelec.fr) by INRIA and R\'egion Lorraine (CPER2007).
\bibliographystyle{elsarticle/elsarticle-num}
|
\section{Introduction}
\label{S:intro}
The evaluation of finite sums involving binomial coefficients appears
throughout the undergraduate curriculum. Students are often exposed to
the identity
\begin{equation}
\sum_{k=0}^{n} \binom{n}{k} = 2^{n}.
\label{two-bin}
\end{equation}
\noindent
Elementary proofs abound: simply choose $x=y=1$ in the
binomial expansion of $(x+y)^{n}$.
The reader is surely aware of many other proofs,
including some combinatorial in nature.
At the end of the previous century, the evaluation of these sums was
trivialized by
the work of H. Wilf and D. Zeilberger \cite{aequalsb}. In the preface to
the charming book \cite{aequalsb}, the authors begin with the phrase
\begin{center}
\texttt{You've been up all night working on your new theory, you found the
answer, and it is in the form that involves factorials, binomial coefficients,
and so on, ...}
\end{center}
\noindent
and then proceed to introduce the method of {\em creative
telescoping}. This technique provides
an automatic tool for the verification of this type of identities.
Even in the presence of a powerful technique, such as the WZ-method, it is
often a good pedagogical idea to present a
simple identity from many different points of view. The reader will
find in \cite{amdeberhan-2012a} this approach with the example
\begin{equation}
\sum_{k=0}^{m} 2^{-2k} \binom{2k}{k} \binom{2m-k}{m} =
\sum_{k=0}^{m} 2^{-2k} \binom{2k}{k} \binom{2m+1}{2k}.
\label{pretty-1}
\end{equation}
The current paper presents
probabilistic arguments for the evaluation of certain
binomial sums. The background required is minimal. The
continuous random variables
$X$ considered here have a probability density
function. This is a nonnegative function
$f_{X}(x)$, such that
\begin{equation}
\Pr(X< x) = \int_{-\infty}^{x} f_{X}(y) \, dy.
\end{equation}
\noindent
In particular, $f_{X}$ must have total mass $1$.
Thus, all computations are reduced to the evaluation of integrals. For
instance, the expectation of a function of the random variable $X$ is
computed as
\begin{equation}
\mathbb{E} g(X) = \int_{-\infty}^{\infty} g(y) f_{X}(y) \, dy.
\end{equation}
\noindent
In elementary courses, the reader has been exposed to normal
random variables, written as
$X \sim N(0,1)$, with density
\begin{equation}
f_{X}(x) = \frac{1}{\sqrt{2 \pi}} e^{-x^{2}/2},
\end{equation}
and exponential random variables, with probability density function
\begin{equation}
f(x;\lambda) = \begin{cases}
\lambda e^{- \lambda x} & \text{ for } x \geq 0; \\
0 & \text{ otherwise.}
\end{cases}
\end{equation}
The examples employed in the arguments presented here have
a gamma distribution with shape parameter $k$ and scale parameter $\theta$,
written as $X \sim \Gamma(k, \theta)$. These
are defined by the density function
\begin{equation}
f(x;k, \theta) =
\begin{cases}
x^{k-1} e^{-x/\theta}/\theta^{k} \Gamma(k), & \quad
\text{ for } x \geq 0; \\
0 & \quad \text{ otherwise}.
\end{cases}
\end{equation}
\noindent
Here $\Gamma(s)$ is the classical gamma function, defined by
\begin{equation}
\Gamma(s) = \int_{0}^{\infty} x^{s-1}e^{-x} \, dx
\end{equation}
\noindent
for $\mathop{\rm Re}\nolimits{s} > 0$. Observe that if $X \sim \Gamma(a,\theta)$, then
$X = \theta Y$ where $Y \sim \Gamma(a,1)$. Moreover $\mathbb{E} X^{n} =
\theta^{n} (a)_{n}$, where
\begin{equation}
(a)_{n} = \frac{\Gamma(a+n)}{\Gamma(a)} = a(a+1) \cdots (a+n-1)
\end{equation}
\noindent
is the Pochhammer symbol. The
main property of these random variables employed in this paper is the
following: assume
$X_{i} \sim \Gamma(k_{i},\theta)$ are independent, then
\begin{equation}
X_{1} + \cdots + X_{n} \sim \Gamma(k_{1} + \cdots + k_{n}, \theta).
\end{equation}
This follows from the fact that
that the density probability function for the sum of two independent
random variables is the convolution of the individual ones.
Related random variables include those with a beta distribution
\begin{equation}
f_{a,b}(x) =
\begin{cases}
x^{a-1}(1-x)^{b-1}/B(a,b) & \quad
\text{ for } 0 \leq x \leq 1; \\
0 & \quad \text{ otherwise}.
\end{cases}
\end{equation}
\noindent
Here $B(a,b)$ is the beta function defined by
\begin{equation}
B(a,b) = \int_{0}^{1} x^{a-1}(1-x)^{b-1} \, dx
\end{equation}
\noindent
and also the symmetric beta distributed
random variable $Z_{c}$, with density proportional to
$(1-x^{2})^{c-1}$ for $-1 \leq x \leq 1.$ The first class of random
variables can
be generated as
\begin{equation}
B_{a,b} = \frac{\Gamma_{a}}{\Gamma_{a} + \Gamma_{b}},
\label{fun-1}
\end{equation}
\noindent
where $\Gamma_{a}$ and $\Gamma_{b}$ are independent gamma distributed with
shape parameters $a$ and $b$, respectively and the second type is
distributed as $1 - 2B_{c,c}$, that is,
\begin{equation}
Z_{c} = 1 - \frac{2 \Gamma_{c}}{\Gamma_{c} + \Gamma'_{c}} = \frac{\Gamma_{c} - \Gamma'_{c}}{\Gamma_{c} +
\Gamma'_{c}},
\label{fun-2}
\end{equation}
\noindent
where $\Gamma_{c}$ and $\Gamma'_{c}$ are independent gamma distributed with
shape parameter $c$. A well-known result is that $B_{a,b}$ and
$\Gamma_{a}+\Gamma_{b}$ are independent in \eqref{fun-1}; similarly,
$\Gamma_{c} + \Gamma'_{c}$ and $Z_{c}$ are independent in \eqref{fun-2}.
\section{A sum involving central binomial coefficients}
\label{S:bincoeff}
Many finite sums may be evaluated via the generating function of terms
appearing in them. For instance, a sum of the form
\begin{equation}
S_{2}(n) = \sum_{i+j=n} a_{i}a_{j}
\end{equation}
\noindent
is recognized as the coefficient of $x^{n}$ in the expansion of $f(x)^{2}$,
where
\begin{equation}
f(x) = \sum_{j=0}^{\infty} a_{j}x^{j}
\end{equation}
\noindent
is the generating function of the sequence $\{ a_{i} \}$. Similarly,
\begin{equation}
S_{m}(n) = \sum_{k_{1} + \cdots + k_{m} =n} a_{k_{1}} \cdots a_{k_{m}}
\end{equation}
\noindent
is given by the coefficient of $x^{n}$ in $f(x)^{m}$. The classical example
\begin{equation}
\frac{1}{\sqrt{1-4x}} = \sum_{j=0}^{\infty} \binom{2j}{j} x^{j}
\label{bin-exp1}
\end{equation}
\noindent
gives the sums
\begin{equation}
\sum_{i=0}^{n} \binom{2i}{i} \binom{2n-2i}{n-i} = 4^{n}
\label{mult-identity-2}
\end{equation}
\noindent
and
\begin{equation}
\label{mult-identity}
\sum_{k_{1}+\cdots + k_{m} = n}
\binom{2k_{1}}{k_{1}} \cdots \binom{2k_{m}}{k_{m}} =
\frac{2^{2n}}{n!} \frac{\Gamma( \tfrac{m}{2} + n )}{\Gamma(\tfrac{m}{2})}.
\end{equation}
\noindent
The powers of $(1-4x)^{-1/2}$ are obtained from the binomial expansion
\begin{equation}
(1 - 4x)^{-a} = \sum_{j=0}^{\infty} \frac{(a)_{j}}{j!}(4x)^{j},
\end{equation}
\noindent
where $(a)_{j}$ is the Pochhammer symbol.
\smallskip
The identity \eqref{mult-identity-2} is elementary and there are
many proofs in the
literature. A nice combinatorial proof of \eqref{mult-identity}
appeared in $2006$ in this journal \cite{valerio-2006a}. In a
more recent contribution, G. Chang and
C. Xu \cite{chang-xu-2011a} present a
probabilistic proof of these identities. Their approach is elementary: take
$m$ independent Gamma random variables $X_{i} \sim
\Gamma(\tfrac{1}{2},1)$ and write
\begin{equation}
\mathbb{E} \left( \sum_{i=1}^{m} X_{i} \right)^{n} =
\sum_{k_{1}+\cdots+k_{m} = n} \binom{n}{k_{1}, \cdots, k_{m}}
\mathbb{E}X_{1}^{k_{1}} \cdots
\mathbb{E}X_{m}^{k_{m}}
\label{identity-0}
\end{equation}
where $\mathbb{E}$ denotes the expectation operator.
For each random variable $X_{i}$, the moments are given by
\begin{equation}
\mathbb{E} X_{i}^{k_{i}} = \frac{\Gamma( k_{i} + \tfrac{1}{2})}{\Gamma(
\tfrac{1}{2} )} = 2^{-2k_{i}} \frac{(2k_{i})!}{k_{i}!} =
\frac{k_{i}!}{2^{2k_{i}}} \binom{2k_{i}}{k_{i}},
\label{moments}
\end{equation}
\noindent
using Euler's duplication formula for the gamma function
\begin{equation}
\Gamma(2z) = \frac{1}{\sqrt{\pi}} 2^{2z-1} \Gamma(z) \Gamma(z+ \tfrac{1}{2})
\end{equation}
(see \cite{nist}, $5.5.5$) to obtain the second form. The expression
\begin{equation}
\binom{n}{k_{1}, \cdots, k_{m}} = \frac{n!}{k_{1}! \, k_{2}! \, \cdots \,
k_{m}!}
\end{equation}
\noindent
for the multinomial coefficients shows that the right-hand side of
\eqref{identity-0} is
\begin{equation}
\frac{n!}{2^{2n}} \sum_{k_{1} + \cdots + k_{m}=n}
\binom{2k_{1}}{k_{1}} \cdots \binom{2k_{m}}{k_{m}}.
\end{equation}
\noindent
To evaluate the left-hand side of \eqref{identity-0}, recall that the sum of
$m$ independent $\Gamma \left(\tfrac{1}{2},1 \right)$
has a distribution of $\Gamma(\tfrac{m}{2},1)$.
Therefore, the left-hand side of \eqref{identity-0} is
\begin{equation}
\frac{\Gamma( \tfrac{m}{2} + n )}{\Gamma( \tfrac{m}{2} )}.
\end{equation}
This gives \eqref{mult-identity}. The special case $m=2$ produces
\eqref{mult-identity-2}.
\section{More sums involving central binomial coefficients}
\label{S:second}
The next example deals with the identity
\begin{equation}
\sum_{k=0}^{n} \binom{4k}{2k} \binom{4n-4k}{2n-2k} = 2^{4n-1} +
2^{2n-1} \binom{2n}{n}
\label{sum-000}
\end{equation}
\noindent
that appears as entry $4.2.5.74$ in \cite{brychkov}. The proof presented here
employs the famous dissection technique, first
introduced by Simpson
\cite{simpson-1759} in the simplification of
\begin{equation}
\frac{1}{2} \left( \mathbb{E}(X_{1}+X_{2})^{2n} +
\mathbb{E}(X_{1}-X_{2})^{2n} \right),
\end{equation}
\noindent
where $X_{1}, \, X_{2}$ are independent random variables distributed
as $\Gamma \left( \tfrac{1}{2}, 1 \right)$.
The left-hand side is evaluated by expanding the binomials to obtain
\begin{multline}
\frac{1}{2} ( \mathbb{E}(X_{1}+X_{2})^{2n} +
\mathbb{E}(X_{1}-X_{2})^{2n} ) = \\
\frac{1}{2}\sum_{k=0}^{2n} \binom{2n}{k} \mathbb{E} X_{1}^{k} \,
\mathbb{E} X_{2}^{2n-k} +
\frac{1}{2} \sum_{k=0}^{2n} (-1)^{k} \binom{2n}{k} \mathbb{E} X_{1}^{k} \,
\mathbb{E} X_{2}^{2n-k} \nonumber
\end{multline}
\noindent
This gives
\begin{equation}
\frac{1}{2} ( \mathbb{E}(X_{1}+X_{2})^{2n} +
\mathbb{E}(X_{1}-X_{2})^{2n} )
= \sum_{k=0}^{n} \binom{2n}{2k} \mathbb{E} X_{1}^{2k} \,
\mathbb{E} X_{2}^{2n-2k}. \nonumber
\label{nice-sum1}
\end{equation}
Using \eqref{moments}, this reduces to
\begin{equation}
\frac{1}{2} \left( \mathbb{E}(X_{1}+X_{2})^{2n} +
\mathbb{E}(X_{1}-X_{2})^{2n} \right)
= \frac{(2n)!}{2^{4n}} \sum_{k=0}^{n} \binom{4k}{2k} \binom{4n-4k}{2n-2k}.
\label{nice-sum1a}
\end{equation}
The random variable $X_{1}+X_{2}$ is $\Gamma(1,1)$ distributed, so
\begin{equation}
\mathbb{E} (X_{1}+X_{2})^{2n} = (2n)!,
\end{equation}
\noindent
and the random variable $X_{1}-X_{2}$ is distributed as
$(X_{1}+X_{2})Z_{1/2}$, where $Z_{1/2}$ is independent of $X_{1}+X_{2}$
and has a symmetric beta distribution with density
$f_{Z_{1/2}}(z) = 1/\pi \, \sqrt{1-z^{2}}$. In particular, the even moments are
given by
\begin{equation}
\label{even-mom}
\frac{1}{\pi} \int_{-1}^{1} \frac{z^{2n} \, dz}{\sqrt{1-z^{2}}} =
\frac{1}{2^{2n}} \binom{2n}{n}.
\end{equation}
\noindent
Therefore,
\begin{equation}
\mathbb{E} (X_{1}-X_{2})^{2n} =
\mathbb{E} (X_{1}+X_{2})^{2n} \, \mathbb{E} Z^{2n}_{1/2} =
\frac{(2n)!}{2^{2n}} \binom{2n}{n}.
\end{equation}
\noindent
It follows that
\begin{equation}
\mathbb{E} (X_{1}+X_{2})^{2n} +
\mathbb{E} (X_{1}-X_{2})^{2n} = (2n)! + \frac{(2n)!}{2^{2n}} \binom{2n}{n}.
\label{nice-sum2}
\end{equation}
The evaluations \eqref{nice-sum1} and \eqref{nice-sum2} imply \eqref{sum-000}.
\section{An extension related to Legendre polynomials}
\label{S:extension}
A key point in the evaluation given in the previous section
is the elementary identity
\begin{equation}
\label{reduc-1}
1 + (-1)^{k} = \begin{cases}
2 & \text{ if } k \text{ is even}; \\
0 & \text{ otherwise. }
\end{cases}
\end{equation}
\noindent
This reduces the number of terms in the sum \eqref{nice-sum1} from $2n$ to $n$.
A similar cancellation occurs for any $p \in \mathbb{N}$. Indeed, the natural
extension of \eqref{reduc-1} is given by
\begin{equation}
\label{reduc-2}
\sum_{j=0}^{p-1} \omega^{jr} = \begin{cases}
p & \text{ if } r \equiv 0 \pmod p; \\
0 & \text{ otherwise};
\end{cases}
\end{equation}
\noindent
Here $\omega = e^{2 \pi i /p}$ is a complex $p$-th root of unity.
Observe that \eqref{reduc-2} reduces to \eqref{reduc-1} when $p=2$.
The goal of this section is to discuss the extension of \eqref{sum-000}. The
main result is given in the next theorem. The
Legendre polynomials appearing in the next theorem are defined by
\begin{equation}
P_{n}(x) = \frac{1}{2^{n} \, n!} \left( \frac{d}{dx} \right)^{n}
(x^{2}-1)^{n}.
\label{legen-def}
\end{equation}
\begin{theorem}
\label{thm-leg}
Let $n, \, p$ be positive integers. Then
\begin{equation}
\sum_{k=0}^{n} \binom{2kp}{kp} \binom{2(n-k)p}{(n-k)p} =
\frac{2^{2np}}{p} \sum_{\ell = 0}^{p-1} e^{i \pi \ell n}
P_{np} \left( \cos \left( \frac{\pi \ell}{p} \right) \right).
\end{equation}
\end{theorem}
\begin{proof}
Replace the random variable $X_{1} - X_{2}$ considered in the previous section,
by $X_{1} + WX_{2}$, where
$W$ is a complex random variable with uniform distribution among the
$p$-th roots of unity. That is,
\begin{equation}
\text{Pr} \left\{ W = \omega^{\ell} \right\} = \frac{1}{p}, \quad
\text{ for } 0 \leq \ell \leq p-1.
\end{equation}
\noindent
The identity \eqref{reduc-2} gives
\begin{equation}
\mathbb{E} W^{r} = \begin{cases}
1 & \text{ if } r \equiv 0 \pmod p; \\
0 & \text{ otherwise.}
\end{cases}
\end{equation}
\noindent
This is the cancellation alluded above.
Now proceed as in the previous section to obtain the moments
\begin{eqnarray}
\mathbb{E}(X_{1} + W X_{2})^{np} & = & \sum_{k=0}^{n} \binom{np}{kp}
\mathbb{E} X_{1}^{(n-k)p} \, \mathbb{E} X_{2}^{kp} \label{sum-3}\\
& = & \frac{(np)!}{2^{2np}}
\sum_{k=0}^{n} \binom{2kp}{kp} \binom{2(n-k)p}{(n-k)p}. \nonumber
\end{eqnarray}
A second expression for $\mathbb{E}(X_{1} + W X_{2})^{np}$ employs
an alternative form of
the Legendre polynomial $P_{n}(x)$ defined in \eqref{legen-def}.
\begin{prop}
\label{legen-1}
The Legendre polynomial is given by
\begin{equation}
P_{n}(x) = \frac{1}{n!} \mathbb{E} \left[ (x + \sqrt{x^{2}-1}) X_{1} +
(x - \sqrt{x^{2}-1}) X_{2} \right]^{n},
\end{equation}
where $X_{1}$ and $X_{2}$ are independent
$\Gamma\left(\tfrac{1}{2},1\right)$ random
variables.
\end{prop}
\begin{proof}
The proof is based on characteristic functions. Compute the sum
\begin{multline}
\label{charac-1}
\mathbb{E} e^{t (x + \sqrt{x^{2}-1}) \, X_{1}} \,
\mathbb{E} e^{t (x - \sqrt{x^{2}-1}) \, X_{2}} = \\
\sum_{k=0}^{\infty} \frac{t^{n}}{n!} \,
\mathbb{E} \left[ (x + \sqrt{x^{2}-1} ) \, X_{1} + (x - \sqrt{x^{2}-1})
\, X_{2} \right].
\end{multline}
The moment generating function for a $\Gamma \left( \tfrac{1}{2}, 1
\right)$ random variable is
\begin{equation}
\mathbb{E} e^{t X} = ( 1 - t)^{-1/2}.
\end{equation}
\noindent
This reduces \eqref{charac-1} to
\begin{equation*}
\left( 1 - t ( x + \sqrt{x^{2}-1}) \right)^{-1/2}
\left( 1 - t ( x - \sqrt{x^{2}-1}) \right)^{-1/2} =
(1 - 2tx + t^{2})^{-1/2}
\end{equation*}
\noindent
which is the generating function of the Legendre polynomials.
\end{proof}
\smallskip
\noindent
This concludes the proof of Theorem \ref{thm-leg}.
\end{proof}
\begin{corollary}
Let $x$ be a variable and $\Gamma_{1}, \, \Gamma_{2}$ as before. Then
\begin{equation}
\mathbb{E} (\Gamma_{1} + x^{2} \Gamma_{2})^{n} = n!x^{n}
P_{n} \left( \tfrac{1}{2}(x + x^{-1} \right).
\label{jou-1}
\end{equation}
\end{corollary}
\begin{proof}
This result follows from Proposition \ref{legen-1} and the
change of variables $x \mapsto \tfrac{1}{2}(x+x^{-1})$, known as
the Joukowsky transform.
\end{proof}
Replacing $x$ by $W^{1/2}$ in \eqref{jou-1} and
averaging over the values of $W$ gives the second expression for
$\mathbb{E}( X_{1} + W X_{2})^{np}$. The
proof of Theorem \ref{thm-leg} is complete.
\section{Chu-vandermonde}
\label{S:chu}
The arguments presented here to prove \eqref{mult-identity-2} can be generalized
by replacing the random variables $\Gamma \left( \tfrac{1}{2},1 \right)$ by
two random variables $\Gamma(a_{i},1)$ with shape parameters $a_{1}$ and
$a_{2}$, respectively. The resulting identity is the Chu-Vandermonde theorem.
\begin{theorem}
Let $a_{1}$ and $a_{2}$ be positive real numbers. Then
\begin{equation}
\sum_{k=0}^{n} \frac{(a_{1})_{k}}{k!} \, \frac{(a_{2})_{n-k}}{(n-k)!} =
\frac{(a_{1}+a_{2})_{n}}{n!}.
\end{equation}
\end{theorem}
The reader will find in \cite{andrews3} a more traditional proof. The paper
\cite{zeilberger-1995} describes how to find and prove this identity in
automatic form.
\medskip
Exactly the same argument for \eqref{mult-identity} provides a multivariable
generalization of the Chu-Vandermonde identity.
\begin{theorem}
Let $\{ a_{i} \}_{1 \leq i \leq m}$ be a collection of $m$ positive real
numbers. Then
\begin{equation}
\sum_{k_{1}+\cdots+k_{m} = n}
\frac{(a_{1})_{k_{1}}}{k_{1}!}
\cdots
\frac{(a_{m})_{k_{m}}}{k_{m}!} = \frac{1}{n!}
(a_{1} + \cdots + a_{m})_{n}.
\end{equation}
\end{theorem}
The final stated result presents a generalization of Theorem \ref{thm-leg}.
\begin{theorem}
Let $n, \, p \in \mathbb{N}, a \in \mathbb{R}^{+}$ and $\omega =
e^{i \pi/p}$. Then
\begin{equation}
\label{gegen-1}
\sum_{k=0}^{n} \frac{(a)_{kp}}{(kp)!} \, \frac{(a)_{(n-k)p}}{((n-k)p)!}
z^{2kp} =
\frac{1}{p} \sum_{\ell=0}^{p-1} e^{i \pi \ell n} z^{np}
C_{np}^{(a)} \left( \tfrac{1}{2}( z \omega^{\ell} + z^{-1} \omega^{-\ell} )
\right).
\end{equation}
\noindent
Here $C_{n}^{(a)}(x)$ is the Gegenbauer polynomial of degree $n$ and
parameter $a$.
\end{theorem}
\begin{proof}
Start with the moment representation for the Gegenbauer polynomials
\begin{equation}
\label{mom-geg}
C_{n}^{(a)}(x) = \frac{1}{n!} \mathbb{E}_{U,V}
\left( U (x+\sqrt{x^{2}-1}) + V ( x - \sqrt{x^{2}-1}) \right)^{n}
\end{equation}
\noindent
with $U$ and $V$ independent $\Gamma(a,1)$ random variables. This
representation is proved in the same way as the proof for the
Legendre polynomial, replacing the exponent $-1/2$ by and exponent $-a$. Note
that the Legendre polynomials are Gegenbauer polynomials with
parameter $a = \tfrac{1}{2}$. This result can also be found in Theorem 3 of
\cite{sun-p-2007a}.
\end{proof}
\begin{note}
The value $z=1$ in \eqref{gegen-1} gives
\begin{equation}
\sum_{k=0}^{n} \frac{(a)_{kp}}{(kp)!} \, \frac{(a)_{(n-k)p}}{((n-k)p)!} =
\frac{1}{p} \sum_{\ell=0}^{p-1} e^{i \pi \ell n}
C_{np}^{(a)} \left( \cos \left( \frac{\pi \ell}{p} \right) \right).
\end{equation}
\noindent
This is a generalization of Chu-Vandermonde.
\end{note}
The techniques presented here may be extended to a variety of situations.
Two examples illustrate the type of identities that may be proven. They
involve the Hermite polynomials defined by
\begin{equation}
H_{n}(x) = (-1)^{n} e^{x^{2}} \left( \frac{d}{dx} \right)^{n}
e^{-x^{2}}.
\end{equation}
\begin{theorem}
Let $m \in \mathbb{N}$. The Hermite polynomials satisfy
\begin{equation}
\label{multiH}
\frac{1}{n!} H_{n} \left( \frac{x_{1} + \cdots + x_{m}}{\sqrt{m}} \right)
= m^{-n/2} \sum_{k_{1}+\cdots + k_{m}}
\frac{H_{k_{1}}(x_{1})}{k_{1}!}
\cdots
\frac{H_{k_{m}}(x_{m})}{k_{m}!}.
\end{equation}
\end{theorem}
\begin{proof}
Start with the moment representation for the Hermite polynomials
\begin{equation}
H_{n}(x) = 2^{n} \mathbb{E}(x + i N)^{n},
\end{equation}
\noindent
where $N$ is normal with mean $0$ and variance $\tfrac{1}{2}$. The details
are left to the reader.
\end{proof}
The moment representation for the Gegenbauer polynomials \eqref{mom-geg}
yields the final result presented here.
\begin{theorem}
Let $m \in \mathbb{N}$. The Gegenbauer polynomials $C_{n}^{(a)}(x)$
satisfy
\begin{equation}
\label{multiG}
C_{n}^{(a_{1}+\cdots+a_{m})}(x) =
\sum_{k_{1}+\cdots + k_{m}=n}
C_{k_{1}}^{(a_{1})}(x)
\cdots
C_{k_{m}}^{(a_{m})}(x).
\end{equation}
\end{theorem}
\begin{remark}
A relation between Gegenbauer and Hermite polynomials is given by
\begin{equation}
\lim\limits_{a \to \infty} \frac{1}{a^{n/2}} C_{n}^{(a)}
\left( \frac{x}{\sqrt{a}} \right) = \frac{1}{n!} H_{n}(x).
\end{equation}
This relation allows to recover easily identity \eqref{multiH}
from identity \eqref{multiG}.
\end{remark}
\medskip
The examples presented here, show that many of the classical identities
for special functions may be established by probabilistic methods. The
reader is encouraged to try this method in his/her favorite identity.
\bigskip
\noindent
\textbf{Acknowledgements}. The work of the second author was
partially supported by NSF-DMS 0070567.
|
\section{Introduction}
MiMeS\footnote{http://www.physics.queensu.ca/$\sim$wade/mimes} (Magnetism in Massive Stars) is a large collaboration that aims to address many issues concerning the magnetism of massive stars. One goal in particular is to determine the global magnetic properties of massive stars with the help of Large Programs (LP) that have been allocated on the high-efficiency high-resolution spectropolarimeters ESPaDOnS (Canada France Hawaii Telescope, Hawaii) and Narval (Telescope Bernard Lyot, France). These programs aim to observe about 200 massive OB field stars (the Survey Component or SC), in order to search for magnetic fields, { confirm those previously suspected,} and derive statistical properties { \citep{wade09,grunhut11b}}. They also aim to observe intensely about 30 already known magnetic massive stars (the Targeted Component { or} TC) in order to map in detail their surface magnetic fields. This smaller sample of stars is dedicated to the study of the interplay of magnetic fields with the stellar structure, environment and evolution at high mass \citep[e.g.][]{grunhut09,oksala11}.
In 2010, the polarimeter HARPSpol was commissioned at the 3.6m-ESO telescope (La Silla, Chile). For the first time, we can access the Southern hemisphere with data quality similar to ESPaDOnS and Narval. Therefore, a Large Program was established to complete the ESPaDOnS/Narval field sample, and to take the first steps toward observing massive stars in various open clusters { and associations} of different ages, to investigate the magnetic field evolution, and the impact of magnetic fields on stellar evolution.
The HARPSpol sample is divided in two components (SC and TC) to follow the same strategy as the Narval and ESPaDOnS LPs. The HARPSpol SC sample contains about 180 stars including $\sim$110 stars in 7 clusters, and $\sim$70 stars in the field of the Galaxy. The former have been selected from the Catalogue of Open Cluster Data \citep{kharchenko05}, while the latter have been chosen from the International Ultraviolet Explorer (IUE) data archive but also from other catalogues or publications containing highly probable magnetic stars, in accordance with the ESPaDOnS/Narval target selection \citep[][]{wade09}.
This HARPSpol Large Program { was} allocated four separate runs over two years. During the first run in May 2011, 57 stars were observed including one magnetic calibrator and one TC target for which the results will be presented in a forthcoming paper. In this letter, we report on the first discoveries of magnetic fields in massive stars with HARPSpol in HD~130807 and HD~122451, and we confirm the magnetic field in HD~105382 previously detected with the low-resolution spectropolarimeter FORS 1 \citep{kochukhov06,hubrig06}. Among the 55 SC stars observed during this run, those three stars are the only ones in which a magnetic field was detected. In Section 2, we present the observations and reduction techniques. In Section 3, we detail the HARPSpol results on each star, and discuss them in Section 4.
\section{Observations}
\begin{table}[t]
\caption{Log of observations. Columns 1 and 2 give the date, Universal Time (UT) and Heliocentric Julian Date of the observations. Columns 3 and 4 give the total exposure time and the number of polarimetric sequences. Columns 5 and 6 give the peak S/N per CCD pixel (at $\sim$501~nm for HD~130807 and at $\sim$518~nm for HD~122451 and HD~105382) in the spectra, and the S/N per 1.4~km.s$^{-1}\;$ (for HD 130807 and HD 105382) and 4.2~km.s$^{-1}\;$ (for HD 122451) pixels in the LSD Stokes $V$ profiles. Columns 7, 8 and 9 give the longitudinal magnetic field, the magnetic detection probability, and the detection type (see text).}
\label{tab:log}
\centering
\begin{tabular}{@{}l@{}c@{}c@{\,\,}c@{\,\,}r@{\,\,}r@{\,\,}r@{$\pm$}l@{\,\,}c@{\,\,}l@{}}
\hline\hline
Date (d/m) & HJD & $t_{\rm exp}$ & \# & S/N & S/N & \multicolumn{2}{c}{$B_{\ell}$} & $P_{\rm det}$ & \\
UT & (2 455 000+) & (s) & & & (LSD) & \multicolumn{2}{c}{(G)} & & \\
\hline
\multicolumn{9}{c}{ HD 130807} \\
23/05 05:23 & 704.724 & 4000 & 1 & 500 & 4600 & 292 & 26 & 1.00000 & DD \\
27/05 06:06 & 708.754 & 1200 & 1 & 520 & 4800 & -94 & 26 & 1.00000 & DD \\
28/05 05:58 & 709.748 & 6000 & 2 & 630 & 5700 & 677 & 21 & 1.00000 & DD \\
\\[-5pt]
\multicolumn{9}{c}{ HD 122451} \\
23/05 04:16 & 704.677 & 1340 & 5 & 1680 & 33200 & -43 & 20\tablefootmark{ a} & 1.00000 & DD \\
27/05 05:21 & 708.723 & 1200 & 20 & 3000 & 61800 & -83 & 14\tablefootmark{ a} & 1.00000 & DD \\
28/05 04:34 & 709.690 & 1680 & 14 & 1260 & 26000 & -66 & 29\tablefootmark{ a} & 0.98173 & ND\\
\\[-5pt]
\multicolumn{9}{c}{ HD 105382} \\
25/05 01:19 & 706.555 & 3200 & 1 & 990 & 9650 & -622 & 26 & 1.00000 & DD \\
26/05 01:59 & 707.582 & 4000 & 1 & 900 & 8840 & -298 & 32 & 1.00000 & DD \\
28/05 00:10 & 709.506 & 4800 & 1 & 920 & 8900 & -406 & 32 & 1.00000 & DD \\
\hline
\end{tabular}
\tablefoot{
\tablefoottext{ a}{ Those values are for HD 122451 B only.}
}
\end{table}
We used the HARPSpol polarimeter \citep{piskunov11}, combined with the HARPS spectrograph \citep{mayor03}, installed at the 3.6m ESO telescope at La Silla Observatory (Chile), yielding spectra with resolving power $\lambda / \Delta \lambda$ of about $105\,000$, and covering the 380--690~nm wavelength region. All spectra were recorded as sequences of 4 individual sub-exposures taken in different configurations of the polarimeter, in order to yield a full circular polarisation analysis, as described by \citet{donati97}. The data were reduced using the package ``REDUCE'' described by \citet{piskunov02}. After reduction, we obtained the intensity Stokes $I$ and the circular polarisation Stokes $V$ spectra of the stars, both normalised to the continuum. A null spectrum ($N$) was also computed in order to diagnose spurious polarisation signatures, and to help to verify that the signatures in the Stokes $V$ spectrum are of stellar origin. The log of the observations is presented in Table 1.
To increase the effective signal to noise ratio (S/N) of our data, we applied the Least Squares Deconvolution \citep[LSD;][]{donati97} procedure using tailored line masks of appropriate temperature and gravity for each star. The masks were first computed using Kurucz ATLAS 9 models of solar abundance \citep{kurucz93}, with intrinsic line depths larger than 0.1. We then excluded from these masks { hydrogen} Balmer lines, and lines whose Land\'e factor is unknown. Finally we have modified the line depths { to} take into account the relative depth of the lines of the observed
spectra. The resulting masks contain 394, 592, and 394 lines for HD 1030807, HD 122451 and HD 105382, respectively. The S/N of the LSD Stokes $V$ profiles is about 10 times larger than the S/N in the original spectra (Table \ref{tab:log}).
In order to perform a reliable magnetic field diagnosis, we have computed the detection probability inside the LSD $V$ profiles \citep[as described in ][]{donati97}. We consider that an observation displays a ``definite detection" (DD) { of Stokes $V$ Zeeman signature} if the probability is larger than 0.99999, a ``marginal detection" (MD) if it falls between 0.999 and 0.99999, and a ``null detection" (ND) otherwise (see Table 1). All observations of HD 130807 and HD 105382 display DD while two DD and one ND have been obtained for HD 122451. The LSD $I$, $V$, and $N$ profiles are plotted in Fig. \ref{fig:lsd}. In almost all of our observations Zeeman signatures, as broad as the $I$ profiles, are clearly detected in the $V$ profiles, while the $N$ profiles are consistent with the noise. These results allow us to confidently affirm that magnetic fields are present at the surface of these stars.
We measured for each observation the line-of-sight component of the magnetic field averaged over the visible stellar surface (the so-called longitudinal magnetic field or $B_{\ell}$), by integrating the $I$ and $V$ profiles over the ranges $[-50,60]$, $[-80,100]$, and $[-70,105]$ km.s$^{-1}$ for HD 130807, HD 122451, and HD 105382, respectively \citep[as described by ][]{alecian09}. The values are reported in Table 1.
\section{Results}
\subsection{HD 130807}
\begin{figure*}
\centering
\includegraphics[width=6cm,clip=true]{meanIVN22.ps}
\includegraphics[width=6cm,clip=true]{meanIVN21.ps}
\includegraphics[width=6cm,clip=true]{meanIVN20.ps}
\caption{LSD $V$ (top), $N$ (middle), and $I$ (bottom) profiles of HD~130807 ({\it left}), HD~122451 ({\it middle}), and HD~105382 ({\it right}). The mean error bars are plotted next to each profile. The $V$ and $N$ profiles have been shifted and amplified for display purpose. The dotted vertical lines indicate the integration ranges for the calculation of $B_{\ell}$. {\it left}: full black line: 28/05, red-dashed line: 27/05, green dot-dashed line: 23/05. {\it middle}: full black line: 27/05, red-dashed line: 23/05, green dot-dashed line: 28/05. {\it right}: full black line: 25/05, red-dashed line: 28/05, green dot-dashed line: 26/05.}
\label{fig:lsd}%
\end{figure*}
{ HD 130807 ($o$ Lup) is member of the Sco-Cen association \citep{kharchenko05}. A companion was detected at an angular distance varying from 0.07 to 0.14 arcsec \citep{perryman97,mcalister90}.
According to the angular separation, the light of both components entered the HARPSpol fibers during our observations of this target.}
{ From a visual inspection} we find that most of the spectrum of HD 130807 is consistent with a synthetic spectrum of a single star of { effective temperature} $T_{\rm eff}=18000$~K, { surface gravity} $\log g=4.25$ (cgs), broadened by $v\sin i=25$~km.s$^{-1}$, calculated using TLUSTY non-LTE atmosphere models and the SYNSPEC code \citep{hubeny88,hubeny92}. However, we observe that all He~{\sc i} lines are substantially weaker than the synthetic ones calculated with solar abundance (Fig. \ref{fig:sphd130807}), while the Si~{\sc ii} lines are considerably stronger. The Si, N and Fe lines show variability in depth and shape on a timescale of 1~d. These characteristics suggest that HD~130807 is an He-weak star with abundance spots on its surface \citep{jaschek74}.
Magnetic signatures are detected in almost all the lines of the spectrum, similar to the LSD one (Fig. \ref{fig:lsd} {\it left}). Many additional lines are observed in the spectrum that could be due to Fe~{\sc ii}, Fe~{\sc iii}, or Ti~{\sc ii} enhancements. All these lines show Zeeman signatures similar to the others, with the same variations from one night to the other. They can therefore be attributed to the same star, rather than a companion, and they are probably the result of the chemical peculiarities at the surface of the star.
A significant shift in radial velocity ($\sim 6$ km/s) is detected in the strongest spectral lines including Balmer lines, between May 22 and May 26-27. The maximum reported angular separation between both visual components implies a distance $\ge17$~AU, and therefore a period $\ge27$ years. This radial velocity shift cannot therefore be due to the reported visual companion. A third companion very close to the primary could explain these variations, but more observations are required to fully understand all the peculiarities observed in the spectrum.
The variations observed in the $V$ profiles over 6 days (Fig. \ref{fig:lsd} {\it left}) can be understood in terms of the oblique rotator (OR) model that consists of an inclined dipole placed inside a rotating star \citep{stibbs50}. The rotational modulation of the shape of the $V$ profiles and of the $B_{\ell}$ values that vary from { -94 to 677~G} (Table 1) suggest that the rotation period of the star should be between 1 and 6 days. More observations, well sampled over the rotation period, are required in order to fully characterise the magnetic field and better constrain the period of HD 130807.
\subsection{HD 122451}
{ HD 122451 ($\beta$ Cen) is a double-lined spectroscopic binary with components of similar effective temperatures (25000 K) and gravities ($\log g=3.5$, cgs), and a $\beta$~Cep-type pulsating primary. The system is highly eccentric ($e = 0.835$) and orbits with a period of 357 days \citep{ausseloos02,ausseloos06,davis05}.}
We obtained { three} observations of HD~122451 during 3 different nights. In order to avoid potential false magnetic detections due to pulsations, we have split each observation into many sequences of four spectra (Table \ref{tab:log}), so that the exposure time for one sequence is much shorter than the pulsation period.
The spectra of HD 122451 clearly show two components of similar temperatures but different broadening, confirming the SB2 nature of the system. We adopt the same definition as \citet{ausseloos06} for the primary and secondary, i.e. as the broad-line and narrow-line components, respectively. In order to measure the $v\sin i\;$ of both stars, we performed a least-squares fit to few individual spectral lines with the sum of two functions calculated as the convolution of a Gaussian of instrumental width and a rotation function as described by \citet{gray92} \citep[see details of the fitting procedure { used by}][]{alecian08}. We find a $v\sin i\;$ of $190\pm20$~km.s$^{-1}\;$ and $75\pm15$~km.s$^{-1}\;$ for the primary and secondary, respectively. When compared with TLUSTY/SYNSPEC synthetic spectra our observations are consistent with $T_{\rm eff}=25000$~K and $\log g=3.5$~(cgs), in agreement with the work of \citet{ausseloos06}. The spectral lines appear distorted and show rapid variations very likely due to $\beta$~Cep-type pulsations. No obvious abundance peculiarity, nor manifestation of circumstellar matter is observed within the spectra.
In Fig. \ref{fig:lsd} ({\it middle}), we superimposed the LSD $I$, $V$, and $N$ profiles of our observations. According to the ephemeris of \citet{ausseloos06}, the 3 observations are roughly at the same orbital phase ($\sim$0.5), and both components have { similar} radial velocities ($\sim$14 and $\sim$4~km.s$^{-1}\;$ for the primary and secondary respectively), which explains why it is difficult to distinguish both components in the profiles. The shape of the LSD $I$ profile shows variations during the run that can be understood in terms of radial pulsations in the primary, which would occasionally broaden the profile. As a result, both components can be clearly distinguished in the profile of May 27 (full black line in Fig. \ref{fig:lsd} {\it middle}), while it is less obvious in the other observations.
\begin{figure}
\centering
\includegraphics[width=9cm,clip=true]{obssyn_hd130807_psp01.ps}
\caption{Spectrum (full black) of HD 130807 plotted around He~{\sc i}~4471~\AA\ and Mg~{\sc ii} 4481 \AA. Synthetic spectra of 18000~K (dashed red) and 17000~K (dot-dot-dot-dashed blue) are overplotted.}
\label{fig:sphd130807}%
\end{figure}
Zeeman signatures are detected in many individual spectral lines, { as in} the LSD $V$ profiles. The signatures are as broad as the secondary profile, meaning that the magnetic field is detected only in the secondary component of the system. However, considering the faint Zeeman signatures in the secondary, and the broad line shape of the primary, a magnetic field of the same strength as the secondary's could exist in the primary, without being detected in our observations. { In order to estimate the $B_{\ell}$ values of the secondary, we need to extract the $V$ and $I$ profiles of the secondary only, from the profiles of the binary. Without any evidence of a magnetic field in the primary, we neglect its contribution to the $V$ profile, which we consider to be entirely from the secondary. On the contrary, the $I$ profile needs to be corrected. With} this aim, we have first fitted the $I$ profile of the binary with the method described above for the individual spectral lines. Then we have subtracted from the observed $I$ profile the fit of the primary. Finally we have measured $B_{\ell}$ using the corrected $I$ and the original $V$ profiles. The values are reported in Table 1.
\subsection{HD 105382}
HD 105382 (= HR 4618) is member of the Sco-Cen association \citep{kharchenko05}. \citet{briquet04} classified it as He-weak with He patches enhanced where Si is depleted, and derived a $T_{\rm eff}$ of $17400\pm800$~K, a $\log g=4.18\pm0.20$ (cgs), a rotation period of $1.295 \pm 0.001$~d and an inclination angle { of the rotation axis to the line of sight} $i=50\pm10$$^{\circ}$.
In our { three} spectra, we observe strong variations in the spectral lines, mainly in He~{\sc i}, Si~{\sc ii} and Fe~{\sc iii}, that are due to abundance spots on the stellar surface described by \citet{briquet04}. Clear Zeeman signatures are detected in the metallic and Balmer lines, as well as in the LSD $V$ profiles (Fig. \ref{fig:lsd} {\it right}). The rotation phases of our observations, calculated with a rotation period of 1.295~d, are very different (0.35, 0.14, and 0.63), and yet the $V$ profiles are all similarly negative (Table 1). According to the OR model, this implies that the magnetic obliquity angle (with respect to the rotation axis) cannot be very high ($|\beta| < 40^{\circ}$ if $i = 50^{\circ}$), otherwise the positive magnetic pole would sometimes appear on the visible stellar hemisphere, creating a positive profile at least once during the run.
HD 105382 was independently discovered as magnetic by \citet{kochukhov06} and \citet{hubrig06}. \citet{briquet07} derived the longitudinal field from FORS~1 observations and found values ranging from $-923$~G to $840$~G. Among their four values, the May 2004 { observation} ($840\pm58$~G) is clearly inconsistent with our data as positive values are not expected. \citet{bagnulo11} very carefully re-reduced the same FORS~1 data and { found values consistent with those of Briquet et al. (2007)} except for that observation ({ for which they derived} $B_{\ell}=-29\pm69$~G).
We performed a least-square sinusoidal fit to our $B_{\ell}$ values simultaneously with the Briquet et al. (2007) data and could find a solution only by removing the May 2004 datapoint. We also performed an independent fit using the re-reduced data of \citet{bagnulo11} and found a similar result. In both cases, the derived period is consistent with that of Briquet et al. (2004). The fitted $B_{\ell}$ values are very similar in both cases, { varying} from $-670$~G to $-20$~G, { and} implying a magnetic obliquity of $\sim38$$^{\circ}\;$ and a polar field strength of $\sim2.3$~kG, assuming a dipole field \citep{borra80}.
\section{Discussion}
We report direct detections of magnetic fields in three hot B-type stars (18000 K - 25000 K), among a sample of 55 stars in which we were searching for magnetic fields with HARPSpol. Two of them (HD 122451 and HD 130807) are completely new detections. For the other one - HD 105382 { - this} is the first direct detection of a Zeeman signature. One of the main MiMeS results is the systematic detection of chemical peculiarities at the surface of magnetic hot stars \citep[and conversely, e.g.][]{grunhut11a}. Among the three stars discussed in this paper two are unambiguously He-weak. The third one (HD 122451) belongs to a binary system with a $\beta$~Cep primary, that makes the interpretation of the spectrum and the detection of peculiarities inside spectral lines very difficult. More observations well sampled over the orbital period of the system are required in order to first confirm a magnetic detection in only the secondary, and then disentangle the pulsation and chemical peculiarity effects.
The interplay between radiative and magnetic forces \citep{hunger99} is usually assumed to be at the origin of the over- or under-abundant He spots at the surface of hot magnetic B stars. These spots are very often correlated with the stellar magnetic fields \citep[e.g.][]{veto90}. In the case of HD~105382, \citet{briquet04} found a large He spot at a latitude of 60$^{\circ}$. If our estimate of its magnetic obliquity is confirmed, this spot would be situated close to the South magnetic pole, demonstrating once more the importance of magnetic fields in the formation of chemical spots.
More observations of these magnetic stars are planned within the HARPSpol large program in order to perform detailed mapping of their magnetic fields, and better confront the models and theories of magnetic massive stars.
\begin{acknowledgements}
We wish to thank the Programme National de Physique Stellaire (PNPS) for their support. JHG and GAW acknowledges support from NSERC. M.B. acknowledges the Fund for Scientific Research -- Flanders for a grant for a long stay abroad, and she is a F.R.S.-FNRS Postdoctoral Researcher. RHDT acknowledges support from NSF grant AST-0908688. This research has made use of the SIMBAD database and the VizieR catalogue access tool, operated at CDS, Strasbourg (France), of INES data from the IUE satellite and of NASAs Astrophysics Data System.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction}
The curvature radiation is the type of the bremsstrahlung radiation when
a radiated charged particle moves along the curved trajectory with the
curvature radius $\rho_0$ and its acceleration is orthogonal to the velocity ${\bf v}$.
The cyclotron rotation of a charged particle in the external magnetic field $B$
is the example of this motion when $\rho_0=v_\perp/\omega_{\rm c}$. Here $\omega_{\rm c}$ is the
cyclotron frequency, $\omega_{\rm c}=eB/{m_{e}}c\gamma$, $e$ and $m_{\rm e}$ are the charge and the
mass of a particle, $\gamma$ is the particle Lorentz factor, and
$v_\perp$ is the component of the particle velocity which
is orthogonal to the magnetic field.
Moving along the circular trajectory, a particle radiates at
harmonics of the cyclotron frequency: $\omega=n\omega_{\rm c}$.
This radiation is called as cyclotron radiation for a
nonrelativistic particle and as synchrotron radiation for a
relativistic particle ($\gamma \gg 1$). For the synchrotron
radiation the maximum of the radiated power turns to be at large
numbers of cyclotron harmonics: $n \simeq \gamma^3$. The total
radiation power $I$ also grows with the particle energy,
$I\propto\gamma^2$. Therefore, the synchrotron radiation of
relativistic particles is presented widely in the space radiation
(Ginzburg \& Syrovatskii 1964).
It is necessary to stress that the length of formation of the curvature
radiation, though is larger than the wave length $\lambda$, is much less than curvature
radius $\rho_0$. So, the properties of the curvature radiation do not differ from
that of the synchrotron radiation in which the cyclotron radius is equal to
the local curvature radius $\rho_0$. The frequency of the maximum of the spectral
power, $\omega \simeq c\gamma^3/\rho_0$, and radiation power,
$I=4/3 \, e^2 c\gamma^4/\rho_0^2$, increases with the particle energy. Here the
dependence
on $\gamma$ is stronger than for the synchrotron radiation because the
curvature is fixed, and does not fall with the energy as for the motion in the
constant magnetic field.
The curvature mechanism of radiation is believed to be connected
with the mechanism of the coherent pulsar radio emission. Indeed,
in the region of the open magnetic field lines in the pulsar
magnetosphere there is a relativistic electron-positron plasma
moving with relativistic velocities along curved magnetic field
lines. For the typical values of curvature radius, $\rho_0 \simeq
10^8$ cm, and Lorentz factor of electrons and positrons,
$\gamma\simeq 10^2$, the characteristic frequency of the curvature
radiation is in the radio band. In the magnetosphere, where the
curvature frequency coincides with the plasma frequency,
$\omega_{\rm p}/\gamma^{3/2}$, we can expect
the collective curvature radiation. Here $\omega_{\rm p}=(4\pi e^2n_{\rm e}/m_{\rm e})^{1/2}$ is the usual
plasma frequency, and $n_{\rm e}$ is the plasma number density. In the strong magnetic field
of the pulsar magnetosphere, when the charged particles can move along magnetic field
lines only, the frequency of the plasma oscillations is $\gamma^{-3/2}$ times
less than the usual plasma frequency $\omega_{\rm p}$.
It seems natural to continue the analogy between the curvature
radiation and the cyclotron radiation for the collective
radiation. But there is the essential difference between them,
which does not permit to rewrite formulas of cyclotron plasma
radiation for the curvature radiation replacing the cyclotron
radius by the curvature radius. The matter is that at each point
of plasma in the magnetic field the distribution of particles over
transverse velocities is isotropic. All directions of particle
transverse motion exist, so that the average velocity equals zero.
It is not so for the curvature radiation when all particles have
only one direction of motion along the magnetic field.
For the plasma physics the problem of the collective curvature
radiation is rather complicated since it demands the consideration
of an essentially nonuniform plasma. It does not result from the
change of parameters of magnetic field and plasma in space. These
effects can be taken into account in the local approximation
because the wave length of radiation is much less than the scales
of inhomogeneities. In order not to lose the curvature radiation
we need to include into consideration the turn of the vector of
anisotropy of the particle distribution function, $f({\bf
p})\propto\delta({\bf p}-p_\parallel {\bf B}/B)$, in space. Here
$p_\parallel$ is the longitudinal particle momentum.
Two parameters of the curvature radiation, i.e., the length of formation
$l_{\rm f}=\rho_0/\gamma\simeq\lambda\gamma^2 \gg \lambda$ and the width
of the radiation directivity $\delta\phi\simeq\gamma^{-1}$, connect by the
relation $l_{\rm f}/\rho_0\simeq\delta\phi$. Thus, the particle is in the
synchronism with the wave (i.e., the particle sees the constant wave phase)
along the path on which the wave intensity changes essentially. The value
of $\lambda$ is the wave length of the curvature radiation,
$\lambda\simeq\rho_0/\gamma^3$.
The problem of calculation of the dielectric permittivity in the
geometrical optics approximation for the nonuniform anisotropic
plasma particle distribution was solved by Beskin, Gurevich and
Istomin (below BGI, 1993). They also described the collective
curvature-plasma interaction, when the electromagnetic waves,
connected with the curvature radiation, are amplified
simultaneously by the Cherenkov mechanism. This effect is absent
in the vacuum. However, this procedure is rather complicated and
demands clear understanding. Because of that there are some
incorrect statements in the literature (see, e.g., Nambu 1989;
Machabeli 1991, 1995).
Apart, another way of investigation of the problem of the
collective curvature radiation was carried out during many years
(Asseo et al. 1983; Larroche \& Pellat 1987; Lyutikov et al. 1999;
Kaganovich \& Lyubarsky 2010). They considered more simple task
connected with the pure cylindrical geometry which can be solved
"exactly". In such a statement the magnetic field lines are
considered to be concentric, the relativistic plasma moving (i.e.,
rotating) along the magnetic field lines owing to the centrifugal
drift directed parallel to the cylindrical axis ($z$-coordinate)
with the velocity $u = c\rho_c/\rho_0 \ll c$. Here again
$\rho_c=c/\omega_{\rm c}$. But this approach cannot be used when
analyse the curvature radiation (Beskin, Gurevich \& Istomin
1988).
Indeed, let us choose the electromagnetic fields of the wave, as was done in all the papers
mentioned above, in the form
\begin{equation}
\left({\bf E},{\bf B}\right)=\left({\bf E}(\rho),{\bf B}(\rho)\right)\times\exp\left\{-i
\omega t+is\phi+ik_z z\right\}.
\end{equation}
Here $\omega$ is the wave frequency, $s$ is integer number
defining the azimuthal wave vector $k_\phi$, and $k_z$ is the
longitudinal wave vector along the cylinder. In this approach the
wave amplitudes ${\bf E}(\rho), {\bf B}(\rho)$ are to be
considered as functions of the radial distance $\rho$ only.
Moreover, not vectors ${\bf E}$ and ${\bf B}$, but their
cylindrical components $(E,B)_\rho, (E,B)_\phi$ and $(E,B)_z$
depend on the coordinate $\rho$ only. It means that the wave
polarization follows the magnetic field, turning from one point
$\phi$ to another. It can be so if we have the definite boundary
condition, e.g., putting the system into the metallic coat. Under
such suggestions we come to the one dimensional problem, which can
be easily solved. Here we will show that such a wave does not have
any relation to the curvature radiation.
Really, let us consider the particle moving exactly along the
circle of radius $\rho_0$ with the constant velocity $v$; this
motion corresponds to the infinite magnetic field. Then the
radiated power is equal to the work of the wave electric field
under the particle electric current. The electric current is
\begin{equation}
{\bf j}= e v \delta(\phi-\Omega t)\, \delta(z)
\frac{\delta(\rho-\rho_0)}{\rho}\, {\bf e}_\phi,
\end{equation}
where $\Omega=v/\rho_0$, and for selected polarization we get
\begin{equation}
\int {\bf jE}{\rm d}{\bf r}=evE_\phi(\rho_0)\exp\left\{-i\omega t+is\Omega t\right\}.
\end{equation}
As we see, the radiation is possible only if $\omega-s\Omega=0$,
i.e., $\omega=k_\phi v$. It is just the condition of Cherenkov,
not curvature radiation. The point is that the wave with such
polarization can not be radiated by the curvature mechanism. The
difference between the curvature wave and the Cherenkov wave is in
the finite interaction time of the bremsstrahlung radiation with a
radiated particle. The freely propagating wave with almost
constant polarization deflects from the direction of a particle
motion. As a result, the nonzero projection of the wave electric
field on the particle velocity (i.e., on the direction of the
electric current) occurs, and the wave takes away the energy from
the particle. This continues the finite time $\tau=l_f/v$ that can
be determined from the relation $\tau(\omega-{\bf kv})\simeq 1$.
For the relativistic particle ($v\simeq c$) $\tau=(\rho_0^2/\omega
c^2)^{1/3}\simeq \rho_0/c\gamma$. Below we will find the real
polarization of the curvature radiation.
The paper is organized as follows. In section 2 we will find that the
polarization of the curvature wave does not correspond to one cylindrical
harmonic. In section 3 it is shown that the nonlinear wave interaction
can lead to significant changes in cylindrical modes propagation.
In section 4 the BGI permittivity tensor will be derived from the permittivity
corresponding to one cylindrical mode. Finally, in section 5 we discuss the
main results of our consideration.
\section{Polarization of the curvature wave}
The radiation field of the electric current density ${\bf j}$ and the electric
charge density $\rho_{\rm e}$ of the moving particle with the charge $e$
is described by the retarded potentials (Landau \& Lifshits 1975):
\begin{eqnarray}
{\bf A} & = & \frac{1}{c}\int\frac{{\bf j}(t')}{R}{\rm d}{\bf r},
\label{A}\\
\Phi & = & \int\frac{\rho_{\rm e}(t')}{R}{\rm d}{\bf r}.
\label{Q}
\end{eqnarray}
Here $t'=t-R/c$ is the retarded time, $R$ is the distance from the charge location
at the time $t'$ to the observer which has the cylindrical coordinates $(\rho, \phi, z)$,
\begin{eqnarray}
R & = & \left[\rho^2+z^2+\rho_0^2-2\rho\rho_0\cos(\phi'-\phi)\right]^{1/2}, \\
\phi' & = & \Omega t'.
\end{eqnarray}
After the Fourier transformation of potentials (\ref{A})--(\ref{Q}) over the time
we obtain
\begin{eqnarray}
{\bf A}_\omega & = & \frac{1}{2\pi}\int{\bf A}(t)\exp\{i\omega t\}{\rm d}t, \\
\Phi_\omega & = & \frac{1}{2\pi}\int\Phi(t)\exp\{i\omega t\}{\rm d}t.
\end{eqnarray}
It is convenient now to replace the integration over time $t$ by the integration over
the retarded time $t'$ and then over the angle $(\phi'-\phi)$. As a result one can obtain
for the cartesian components ($x,y,z$) of the vector potential ${\bf A}$ and the scalar
potential $\Phi$
\begin{equation}
\left[{\bf A}_{\omega};\,\Phi_{\omega}\right]=\frac{e\rho_0}{2\pi c}\exp\{i\omega\phi/\Omega\}
\left[-K_s;\,K_c;\, 0;\,\frac{c}{v}K_0\right].
\label{eqn1}
\end{equation}
Here the quantities $K_0, K_s, K_c$ are the functions of coordinates $\rho$ and $z$ only and
they are equal to
\begin{eqnarray}
K_0 & =& \int\frac{\exp\{i\omega(R/c+\Omega^{-1})\alpha\}}
{R+v\rho\sin\alpha/c}{\rm d}\alpha, \nonumber \\
K_s & = & \int\frac{\exp\{i\omega(R/c+\Omega^{-1})\alpha\}\sin\alpha}
{R+v\rho\sin\alpha/c}{\rm d}\alpha, \\
K_c & =& \int\frac{\exp\{i\omega(R/c+\Omega^{-1})\alpha\}\cos\alpha}
{R+v\rho\sin\alpha/c}{\rm d}\alpha, \nonumber \\
R & = & (\rho^2+z^2+\rho_0^2-2\rho\rho_0\cos\alpha)^{1/2}. \nonumber
\end{eqnarray}
The expression (\ref{eqn1}) is valid at any point ${\bf r}$, i.e., not only in
the wave zone. The dependence over the angle $\phi$ is given by the
exponent $\exp\{i\omega\phi/\Omega\}$. From the periodicity over $\phi$ we
have $\omega=s\Omega$.
The key point of the above expansion (\ref{eqn1}) is that the radiated wave is the
superposition of three harmonics: $s$,$s-1$, and $s+1$. For example, the
azimuthal electric field $E_{\phi\omega}$ is equal to
\begin{eqnarray}
&&E_{\phi\omega} = \frac{i\omega}{v}\left(-\frac{\rho_0}{\rho}\Phi_\omega
+\frac{v}{c}A_{\phi\omega}\right) \nonumber \\
&&= -i\frac{e\rho_0\omega}{2\pi v^2}e^{is\phi}\left[
\frac{\rho_0}{\rho}K_0-\frac{v^2}{c^2}\left(K_s\sin\phi + K_c\cos\phi\right)
\right].
\label{eqn6}
\end{eqnarray}
The first term in Eqn. (\ref{eqn6}), which is proportional to the
scalar potential $\Phi$, is not important in the wave zone, $\rho
\gg \rho_0$, but is significant in the near zone on the particle
trajectory $\rho=\rho_0$. Due to this term, the particle, which is
in the resonance with one of three harmonics, say with $s$
($\omega= s\Omega$), is beaten out of the synchronism by neighbour
harmonics $s\pm 1$. The electric field $E_{\phi\omega}$ changes
its sign during the time $\tau$. The synchronism condition, i.e.,
$1-\cos\Omega\tau\simeq 1-v^2/c^2=\gamma^{-2}$, defines the time
$\tau$,
\begin{equation}
\tau\simeq 1/\Omega\gamma=\rho_0/c\gamma,
\end{equation}
which coincides with the time of formation of the curvature radiation.
Thus, the radiated curvature wave consists of three harmonics $s,
s\pm 1$ with the fixed relation between their amplitudes. Namely,
this circumstance provides the curvature mechanism of the
radiation. Appearance of harmonics $s\pm 1$ except the resonant
one $s=\omega/\Omega$ is due to the additional modulation of the
radiation field induced by a modulation of the particle electric
current having the harmonic $s=1$. Now one can understand why the
simple problem of the collective curvature radiation in the
cylindrical geometry with only one azimuthal harmonic $\exp\{i s
\phi\}$ does not reveal any significant amplification of waves
(Asseo et al., 1983; Lyutikov et al., 1999; Kaganovich \&
Lyubarsky, 2010). In this case the chosen wave polarization does
not contain primordially the curvature mechanism.
\section{Collective triple radiation}
In the previous section it was shown that the curvature radiation
of one charged particle can not be described in the pure
cylindrical geometry by one azimuthal harmonic $\exp\{is\phi\}$.
In a collective radiation the modulation of the particle electric
current appears together with electromagnetic field excitation.
Because of that the resonant azimuthal harmonic
$s=\omega\rho/v_\phi$ mixes with harmonics of the electric current
modulation and produces all possible values of $s$. Further in the
section 4 we will see the all azimuthal harmonics $s$ give
contribution to the response of a media on an electromagnetic
field. But in this section it will be demonstrated that the
collective curvature radiation of only triplex of azimuthal
harmonics $(s, s\pm 1)$ differs significantly from that of one
harmonic $s$ as it is usually considered in the literature.
Let us consider the simple cylindrical one-dimensional problem of
radiation of the cold stream of plasma particles with the charge
$e$ and the mass $m_{\rm e}$ moving along the infinite azimuthal
magnetic field $B_0 = B_{\phi}$. In this case the particles can
move only in $\phi$-direction with the velocity $v_{\phi}$ at
different cylindrical radius $\rho$. The unperturbed particle
density $n^{(0)}$ and velocity $v_\phi^{(0)}$ are constants, i.e.,
they do not depend on $\rho$. The electric current ${\bf j}$ has
only $\phi$-component as well as $B_z$-component of the wave
magnetic field ($B_\rho = B_\phi=0$). Accordingly, the wave
electric field has two components $E_\rho$ and $E_\phi$ ($E_z=0$).
The dependence of the wave fields over time and coordinates is the following
\begin{equation}
[E_{\rho}; \, E_{\phi}; \, B_{z}] = [E_{\rho}(\rho); \, E_{\phi}(\rho); \, B_{z}(\rho)]\exp\{-i\omega t +is\phi\}.
\end{equation}
Then, we obtain from Maxwell equations
\begin{eqnarray}
\frac{{\rm d} E^{(\sigma)}_{\phi}}{{\rm d} \rho} = \frac{i \sigma}{\rho} E^{(\sigma)}_{\rho} -
i \frac{\rho}{\sigma}\frac{\omega^2}{c^2} E^{(\sigma)}_{\rho} - \frac{E^{(\sigma)}_{\phi}}{\rho},\\
\frac{{\rm d} E^{(\sigma)}_{\rho}}{{\rm d} \rho} = - i \frac{\sigma}{\rho} E^{(\sigma)}_{\phi} +
\frac{4 \pi}{\omega}\frac{\sigma}{\rho} j^{(\sigma)}_{\phi} - \frac{E^{(\sigma)}_{\rho}}{\rho}.
\end{eqnarray}
Here index $\sigma$ corresponds to one of three harmonics $s$ or
$s\pm 1$. For simplicity we use here the dimensionless variable
$r$ defined as $r = \rho\omega/c$, as well as quantities $\Lambda
= \omega^2_{p}/(\omega^2\gamma^3)$ and $ J_{\sigma} = 4 \pi
j^{(\sigma)}_{\phi}/(\Lambda \omega)$. Here $\omega_{\rm p}=(4\pi
ne^2/m_{e})^{1/2}$ is plasma frequency, and $\gamma$ is the
Lorentz-factor of the particle motion:
$\gamma=(1-v_\phi^2/c^2)^{-1/2}$. After these definitions the
equations above take the following form
\begin{eqnarray}
\frac{{\rm d} E^{(\sigma)}_{\phi}}{{\rm d} r} & = & \frac{i \sigma}{r} E^{(\sigma)}_{\rho} - i \frac{r}{\sigma}
E^{(\sigma)}_{\rho} - \frac{E^{(\sigma)}_{\phi}}{r},\label{eq78}\\
\frac{{\rm d} E^{(\sigma)}_{\rho}}{{\rm d} r} & = & - i \frac{\sigma}{r} E^{(\sigma)}_{\phi} +
\Lambda \frac{\sigma}{r} J_{\sigma} - \frac{E^{(\sigma)}_{\rho}}{r}.
\end{eqnarray}
As was already stressed, we consider here the interaction of three
waves
$s,s\pm 1$. It is important that they are not independent and their
interaction is realized by the static electric field
$[E_\rho(\rho); \, E_\phi(\rho)] \exp\{i\phi\}$ having the first
azimuthal harmonic $s=1$. This electrostatic field turns to be the
result of nonlinear interactions of high frequency neighbour
harmonics $s$ and $s\pm 1$. Equations for the mode $s = 1$ under
the same definitions are
\begin{eqnarray}
\frac{{\rm d} E_{\phi}}{{\rm d} r} & = & \frac{i}{r} E_{\rho} - \frac{E_{\phi}}{r},\\
\frac{{\rm d} E_{\rho}}{{\rm d} r} & = & - i \frac{1}{r} E_{\phi} + \Lambda Z - \frac{E_{\rho}}{r}.
\label{eq90}
\end{eqnarray}
Here $ Z = 4 \pi n e c/(\Lambda \omega)$.
To determine the response of the stream on the electromagnetic fields of the wave one can use
the continuity and Euler equations
\begin{eqnarray}
&&\frac{\partial n}{\partial t}+{\rm \nabla}(n {{\bf v}})=0,\\
&&\left(\frac{\partial}{\partial t}+{\bf v}\nabla\right){\bf p}
=e\left({\bf E}+\left[\frac{{\bf v}}{c},{\bf B}\right]\right).
\end{eqnarray}
It is easily to understand that only the $\phi$-component of Euler
equation is needed, while the radial component just provides us
the equilibrium configuration across the infinite magnetic field.
We represent the plasma number density and the plasma velocity as
the expansion over powers of the wave amplitude
\begin{eqnarray}
v_{\phi} = v^{(0)}_{\phi} + \delta v^{(1)}_{\phi} + \delta v^{(2)}_{\phi} + ...,\\
n = n^{(0)} + \delta n^{(1)} + \delta n^{(2)} + ....
\end{eqnarray}
The linear response can be easy found
\begin{eqnarray}
n^{(1)} & = & n^{(0)} \frac{k v^{(1)}_{\phi}}{\omega-kv_\phi^{(0)}},
\label{denom1} \\
v^{(1)}_{\phi} & = & i \frac{e E_{\phi}}{m_{\rm e} \gamma^3 (\omega-kv_\phi^{(0)})},
\label{denom2}
\end{eqnarray}
where $k=s/\rho$. On the other hand, for the nonlinear current the nonlinear relation
between $\delta v_{\phi}$ and $\delta p_{\phi}$ should be taken into account
\begin{equation}
\delta p_{\phi} = m_{\rm e}\gamma^3 \delta v_{\phi} -
\frac{3}{2}m_{\rm e} v^{(0)}_{\phi}\gamma^5 \frac{(\delta v_{\phi})^2}{c^2}.
\end{equation}
The result of cumbersome but straightforward calculation is
\begin{widetext}
\begin{eqnarray}
J_s = \frac{1}{1- s v^{(0)}_{\phi}/r}\left[i\frac{E^{s}_{\phi}}
{1- s v^{(0)}_{\phi}/r} + \alpha\frac{r}{v^{(0)}_{\phi}}\left(A_{s,s-1}\frac{E^{s-1}_{\phi}E^{1}_{\phi}}{1- (s -1)
v^{(0)}_{\phi}/r} -
A_{s,s+1}\frac{E^{s+1}_{\phi}E^{1*}_{\phi}}{1- (s +1) v^{(0)}_{\phi}/r}\right)\right],\label{Jeq1}\\
J_{s-1} = \frac{1}{1- (s -1) v^{(0)}_{\phi}/r}\left[i\frac{E^{s-1}_{\phi}}{1- (s -1) v^{(0)}_{\phi}/r} - \alpha\frac{r}{v^{(0)}_{\phi}}A_{s,s-1}\frac{E^{s}_{\phi}E^{1*}_{\phi}}{1- s v^{(0)}_{\phi}/r}\right],\\
J_{s+1} = \frac{1}{1- (s +1) v^{(0)}_{\phi}/r}\left[i\frac{E^{s}_{\phi}}{1- (s +1) v^{(0)}_{\phi}/r} + \alpha\frac{r}{v^{(0)}_{\phi}}A_{s,s+1}\frac{E^{s}_{\phi}E^{1}_{\phi}}{1- s v^{(0)}_{\phi}/r}\right],\\
Z = \frac{1}{\left(v^{(0)}_{\phi}\right)^2} \left[i\frac{E_{1}}{1/r} + \alpha \left(\frac{E^{s+1}_{\phi}E^{s*}_{\phi}}{(1- (s +1) v^{(0)}_{\phi}/r)(1- s v^{(0)}_{\phi}/r)} + \frac{E^{s}_{\phi}E^{(s-1)*}_{\phi}}{(1- s v^{(0)}_{\phi}/r)(1- (s -1) v^{(0)}_{\phi}/r)}\right)\right],\label{Jeq2}\\
A_{i,j} = \frac{1}{1- i v^{(0)}_{\phi}/r} + \frac{1}{1- j v^{(0)}_{\phi}/r} -3 \gamma^2,\nonumber
\end{eqnarray}
\end{widetext}
Here $\alpha = e/(m_{e}c\gamma^3\omega)$ is the particle velocity
divided over the velocity of light. The same quantities for plane
waves can be found in (BGI, 1993). Equations above are evaluated
with vacuum initial condition for the normal mode that can be
presented analytically, $E^{(\sigma)}_{\phi} = -
J^{'}_{\sigma}(r), E^{(\sigma)}_{r} = i {\sigma} J_{\sigma}(r)/r$.
Here $J_\sigma(r)$ is the Bessel function. It should be noted that
the singularity in equations (\ref{eq78}) is passed smoothly by
additional small term $ +i \varepsilon$ in the resonance
denominators in (\ref{denom1})--(\ref{denom2}).
\begin{figure*}
\includegraphics[scale=0.3]{fig1.pdf}
\caption{Model calculations of two cases, $\Lambda = 10^{-2}$, $\nu = 1 GHz$, $\gamma =5$, $s = 125$.}
\label{fig1}
\end{figure*}
In numerical calculations equations (\ref{eq78})--(\ref{eq90}) for $\sigma = s$ and $\sigma = s \pm 1$
were solved with two different values for the quantities $J_{\sigma}$ and $Z$. In the first case
we neglect non-linear terms in (\ref{Jeq1})--(\ref{Jeq2}), while the second one corresponds
to the full non-linear problem. On Fig. ~\ref{fig1} the results obtained for this cases are
presented. For better representation of the influence of the nonlinear current, we choose
the amplitudes of $s - 1$ and $s + 1$ modes twenty times higher than the amplitude
of the $s$ mode. In reality the $s$-mode interacts with the whole continuum of modes,
so this model assumption is rather reasonable. Fig. ~\ref{fig1} shows that in this case
the intensity of the wave $|E|^2$ is approximate 2.5 times larger than in the case when
the nonlinear current is neglected. Hence, one can conclude that three wave interaction
is rather effective.
Thus, we have shown that the triplex of cylindrical harmonics,
which corresponds better to the curvature mechanism, is amplified
more effective than the separated harmonic having the single value
of the azimuthal number. In fact the real polarization of the
collective curvature mode can be obtained only by calculating the
permittivity tensor of the streaming plasma in the strong curved
magnetic field. The solution of wave equations produces not only
the dispersive equation for normal waves, $\omega =\omega({\bf
k})$, but defines also their polarization. A priory it is unclear
what polarization corresponds to unstable modes.
At first sight, the problem considered above is essentially
nonlinear and has no direct connection with the question of the
linear wave amplification. We included nonlinearity only in order
to connect harmonics $s,\, s\pm 1$ self-consistently. And
appearance of neighbour harmonics $s\pm 1$, even for small
nonlinearity, strongly change the $s$-mode amplification. It is
clear also that interaction of $s\pm 1$ waves with the field $s=1$
will result in all azimuthal harmonics.
\section{Tensor derivation}
In this section we will show that the asymptotic behaviour of the BGI dielectric tensor in the
case of large enough curvature radius $\rho_{0}$ can be found directly from the plasma response on
the one cylindrical mode. For the infinite toroidal magnetic field
only the response to the toroidal component of the wave electric field $E_{\phi}$ is to be
included into consideration (Beskin 1999). Here and below we consider the stationary medium
only, so the time dependence can be chosen as $\exp\{- i \omega t\}$. Making summation over
all cylindrical modes, one can write down
\begin{equation}
D_{\phi} (\rho, \phi) = E_{\phi} (\rho, \phi) -
\sum\limits_{s = - \infty}^{\infty} E_{\phi}(\rho, s) K(\rho, s) \exp\{is\phi\},
\label{eqm}
\end{equation}
where
\begin{equation}
K (\rho, s) = \frac{4 \pi e^2}{\omega} \int
\frac{v_{\phi}}{\omega - s v_{\phi}/\rho} \frac{\partial f^{(0)}}{\partial p_{\phi}}
{\rm d}p_{\phi}.
\end{equation}
Here $f^{(0)}(p_{\phi})$ is unperturbed distribution function.
Making the Fourier transformation
\begin{equation}
E_{\phi}(\rho,s) = \frac{1}{2\pi} \int\limits_{0}^{2\pi} E_{\phi}(\rho,\phi') \exp\{- i s\phi'\} {\rm d}\phi'
\end{equation}
and the transition to cartesian coordinate system one can obtain:
\begin{eqnarray}
D_{x} = E_{x} + \frac{1}{2\pi}\int\frac{\rho'{\rm d}\rho'{\rm d}\phi'}{\rho'}
\sum\limits_{s = - \infty}^{\infty}E_{\phi}(\rho', \phi')\delta(\rho - \rho')
\times\nonumber \\
K(\rho, s) \, \exp\{i s (\phi - \phi')\}\sin \phi,\\
D_{y} = E_{y} - \frac{1}{2\pi}\int\frac{\rho'{\rm d}\rho'{\rm d}\phi'}{\rho'}
\sum\limits_{s = - \infty}^{\infty}E_{\phi}(\rho', \phi')\delta(\rho - \rho')
\times \nonumber \\
K(\rho, s) \, \exp\{i s (\phi - \phi')\}\cos \phi.
\end{eqnarray}
We choose the local coordinate system with the $y$-axis directed
along the magnetic field and the $x$-axis, that is orthogonal to
it. From the above equations one can obtain the permittivity
kernel components
\begin{eqnarray}
\varepsilon_{yy}({\bf{r}}, {\bf{r'}}) = 1 - \frac{1}{2\pi}\frac{1}{\rho'}
\sum\limits_{s = - \infty}^{\infty}\delta(\rho - \rho') K(\rho, s)\times
\nonumber \\
\exp\{i s (\phi - \phi')\}\cos \phi \cos \phi';\\
\varepsilon_{yx}({\bf{r}}, {\bf{r'}}) = \frac{1}{2\pi}\frac{1}{\rho'}
\sum\limits_{s = - \infty}^{\infty}\delta(\rho - \rho') K(\rho, s) \times
\nonumber \\
\exp\{i s (\phi - \phi')\}\cos \phi \sin \phi';\\
\varepsilon_{xy}({\bf{r}}, {\bf{r'}}) = \frac{1}{2\pi}\frac{1}{\rho'}
\sum\limits_{s = - \infty}^{\infty}\delta(\rho - \rho') K(\rho, s)\times
\nonumber \\
\exp\{i s (\phi - \phi')\}\sin \phi \cos \phi',\\
\varepsilon_{xx}({\bf{r}}, {\bf{r'}}) = 1 - \frac{1}{2\pi}\frac{1}{\rho'}
\sum\limits_{s = - \infty}^{\infty}\delta(\rho - \rho') K(\rho, s)\times\nonumber \\
\exp\{i s (\phi - \phi')\}\sin \phi \sin \phi',
\end{eqnarray}
that provides the material relationship
\begin{equation}
D_{i}({\bf{r}}) = \int \varepsilon_{ij}({\bf{r}}, {\bf{r'}})E_{j}({\bf{r'}}) {\rm d} {\bf{r'}}.
\end{equation}
It should be noted that the operator presented above satisfies the needed symmetry condition
\begin{equation}
\varepsilon_{ij}({\bf{r}}, {\bf{r'}}, \omega) = \varepsilon_{ji} ({\bf{r'}}, {\bf{r}}, -\omega)
\end{equation}
(it is provided by the condition $K(r, s, \omega) = K(r, -s, -\omega)$). As it is well-known
(Kadomtsev 1965; Bornatici \& Kravtsov 2000), it is this symmetrical form of permittivity tensor
that is to be used for the calculation of components of the permittivity tensor
$\varepsilon_{ij} (\omega, {\bf {k}}, {\bf {r}})$
\begin{equation}
\varepsilon_{ij} (\omega, {\bf {k}}, {\bf {\boldsymbol\eta}} \to {\bf {r}})
= \int \varepsilon_{ij}(\omega, {\bf{\boldsymbol\xi}}, {\bf{\boldsymbol\eta}}) \exp\{-i\bf{k\boldsymbol\xi}\}{\rm d} {\bf{r}} .
\label{kadom}
\end{equation}
Here ${\boldsymbol{ \boldsymbol\eta}} = ({\bf r} + {\bf r'})/2$,
${\bf \boldsymbol\xi} = \bf{r} - \bf{r'}$. It is important that
the above tensor only describes correctly wave-particle
interaction in inhomogeneous media with slowly varying parameters
(Bernstein \& Friedland 1984).
Substituting now the kernel components, one can find
\begin{eqnarray}
\varepsilon_{xx}(\omega, {\bf {k}}, {\bf {\boldsymbol\eta}}) = 1 - \frac{1}{2\pi}\int {\rm d}{\bf \boldsymbol\xi} \exp\{-i{\bf{k \boldsymbol\xi}\}}\frac{1}{|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|}\times\nonumber\\\sum\limits_{s = - \infty}^{\infty}\delta(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|) K(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|, s) \times
\nonumber\\
\exp\{i s (\phi - \phi')\}\sin \phi \sin \phi',
\label{eps1} \\
\varepsilon_{xy}(\omega, {\bf {k}}, {\bf {\boldsymbol\eta}}) = \frac{1}{2\pi}\int {\rm d}{\bf \boldsymbol\xi} \exp\{-i{\bf{k \boldsymbol\xi}\}}\frac{1}{|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|}\times\nonumber\\\sum\limits_{s = - \infty}^{\infty}\delta(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|) K(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|, s)\times\nonumber \\
\exp\{i s (\phi - \phi')\}\sin \phi \cos \phi',
\label{eps12} \\
\varepsilon_{yx}(\omega, {\bf {k}}, {\bf {\boldsymbol\eta}}) = \frac{1}{2\pi}\int {\rm d}{\bf \boldsymbol\xi} \exp\{-i{\bf{k \boldsymbol\xi}\}}\frac{1}{|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|}\times\nonumber\\\sum\limits_{s = - \infty}^{\infty}\delta(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|) K(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|, s)\times\nonumber \\
\exp\{i s (\phi - \phi')\}\cos \phi \sin \phi',
\label{eps21} \\
\varepsilon_{yy}(\omega, {\bf {k}}, {\bf {\boldsymbol\eta}}) = 1 -\frac{1}{2\pi}\int {\rm d}{\bf \boldsymbol\xi} \exp\{-i{\bf{k \boldsymbol\xi}\}}\frac{1}{|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|}\times\nonumber\\\sum\limits_{s = - \infty}^{\infty}\delta(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|) K(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|, s) \times\nonumber \\
\exp\{i s (\phi - \phi')\}\cos \phi \cos \phi'.
\label{eps3}
\end{eqnarray}
In this equations, the angles $\phi$ and $\phi'$ are the functions of polar angles of vectors
$\bf{\boldsymbol\eta}$ and $\bf{\boldsymbol\xi}$, $\alpha_{\eta}$, and $\alpha_{\xi}$
\begin{eqnarray}
\sin \phi = \frac{|\boldsymbol\eta|\sin\alpha_{\eta} + (|\boldsymbol\xi|/2)\sin\alpha_{\xi}}{|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|},\\
\cos \phi' = \frac{|\boldsymbol\eta|\cos\alpha_{\eta} - (|\boldsymbol\xi|/2)\cos\alpha_{\xi}}{|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|}.
\end{eqnarray}
As a result, integrals above are reduced to integration over $\bf{\boldsymbol\xi}$, which is perpendicular
to $\bf{\boldsymbol\eta}$. On the other hand, the expression for the delta-functions in (\ref{eps1})--(\ref{eps3})
is the following:
\begin{eqnarray}
\delta(...) & = & \frac{\delta(\theta - \pi/2)}{(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|)'_{\theta}}
\nonumber \\
& + & \frac{\delta(\theta + \pi/2)}{(|{\bf \boldsymbol\eta + \boldsymbol\xi}/2| -|{\bf \boldsymbol\eta - \boldsymbol\xi}/2|)'_{\theta}},
\label{deltaf}
\end{eqnarray}
where $\theta$ is the angle between vectors $\bf{\boldsymbol\eta}$ and ${\bf{\boldsymbol\xi}}$. So, the integration
over angles can be done easily. Finally, from the transition ${\bf {\boldsymbol\eta}} \to {\bf {r}}$, one
can obtain $\cos\alpha_{\eta} \to \cos\alpha_{r} =1$. Hence, according to (\ref{deltaf})
$({\bf k \boldsymbol\xi}) = k_{\parallel} |\boldsymbol\xi|$, where $k_{\parallel}$ is the component of the wave vector
parallel to external magnetic field.
The property of the absence of $k_{\perp}$ is very important, it provides the same symmetry as
it was in the case of homogeneous medium: $\varepsilon_{ij}(- \omega, -{\bf k}, -{\bf B}, {\bf r}) =
\varepsilon_{ji}(\omega, {\bf k}, {\bf B}, {\bf r})$ (Istomin 1994). This result differs from
one obtained by Lyutikov at al. (1999). In this work the importance of transformation (\ref{kadom})
is neglected.
Using finally the Taylor expansion over $|\boldsymbol\xi|$ and the reduction of resonant denominator to delta-function,
one can obtain:
\begin{eqnarray}
&& \sum(...)\frac{1}{\omega|{\bf \boldsymbol\eta + \boldsymbol\xi}/2|/v_{\phi} - s}
\nonumber \\
&& \to i\pi \int (...) \delta \left[s - \frac{\omega(|\boldsymbol\eta|^2 + |\boldsymbol\xi|^2/4)^{1/2}}{v_{\phi}}\right]{\rm d}s.
\end{eqnarray}
As a result, one can write down
\begin{eqnarray}
\varepsilon_{xx} & = & -i\frac{8\pi^2 e^2}{\omega}\int F''(\kappa) \frac{v_{\phi}}{\omega}\frac{\partial f^{(0)}}{\partial p_{\phi}}{\rm d} p_{\phi},\\
\varepsilon_{xy} & = & - \varepsilon_{yx} = \frac{8\pi^2 e^2}{\omega}\int F'(\kappa) \frac{\rho^{1/3}_0 v^{2/3}_{\phi}}{\omega^{2/3}}\frac{\partial f^{(0)}}{\partial p_{\phi}}{\rm d} p_{\phi},\\
\varepsilon_{yy} & = & -i\frac{8\pi^2 e^2}{\omega}\int F(\kappa) \frac{\rho^{2/3}_0 v^{ 1/3}_{\phi}}{\omega^{1/3}}\frac{\partial f^{(0)}}{\partial p_{\phi}}{\rm d} p_{\phi}.
\end{eqnarray}
Here
\begin{eqnarray}
F(\kappa) & = & \frac{1}{\pi}\int\limits^{+\infty}_{0} \exp\{i\kappa t + it^3/3\} {\rm d} t,\\
\kappa & = & \frac{2(\omega - k_{\parallel} v_{\phi})}{\omega^{1/3} v^{2/3}_{\phi}}\rho^{2/3}_{0},
\end{eqnarray}
prime means the derivative, and $\rho_0$ is the curvature radius of magnetic field.
Due to high enough curvature radius of field lines in the pulsar
magnetosphere, one can use the asymptotic behaviour of $F(\kappa)$
for $\kappa \gg 1$
\begin{equation}
F(\kappa) \approx \frac{i}{\pi \kappa} + \frac{2i}{\pi \kappa^4} + ...
\end{equation}
After integration by parts, the final result is
\begin{equation}
\varepsilon_{ij} =
\begin{pmatrix}
1 - \frac{3}{2}\left<\frac{\omega^2_{pl} v^2_{\parallel}}{\gamma^3 \rho^2_0 \tilde{\omega}^4}\right> && - i\left<\frac{\omega^2_{pl} v_{\parallel}}{\gamma^3 \rho_0 \tilde{\omega}^3}\right> \cr
i\left<\frac{\omega^2_{pl} v_{\parallel}}{\gamma^3 \rho_0 \tilde{\omega}^3}
\right>&& 1 - \left<\frac{\omega^2_{pl}}{\gamma^3 \tilde{\omega}^2}\right>
\cr
\end{pmatrix}
\label{BGItensor}
\end{equation}
Here by definition
$\tilde{\omega} = \omega - {\bf k v}$,
and the brackets $<>$ denote both the averaging over the particle distribution
function $f_{e^{+}, e^{-}}(p_{\phi})$ and the summation over the types of particles:
\begin{equation}
<(...)> \, = \sum_{e^{+}e^{-}} \int (...) f^{(0)}_{e^{+},e^{-}}(p_{\phi}){\rm d} p_{\phi}.
\end{equation}
We see that the tensor above is just the BGI tensor, that leads to
instability of the so-called curvature plasma modes. In the limit
$\rho_{0} = \infty$ this tensor, as expected, tends to the
dielectric permittivity of a homogeneous plasma. The nonzero
components $\varepsilon_{xy}, \varepsilon_{yx}$ and
$\delta\varepsilon_{xx}=\varepsilon_{xx}-1$ in the tensor
$\varepsilon_{ij}$ (\ref{BGItensor}) for the finite curvature are
due to nonlocal properties of the plasma response on the
electromagnetic wave in curved magnetic field. The parameter of
nonlocality $(v_\parallel/\tilde{\omega})/\rho_0$ is the ratio of
the formation length of radiation to the curvature radius. For
vacuum $\tilde{\omega}\simeq\omega/\gamma^2$, and the length
$v_\parallel/\tilde{\omega}$ coincides with the length of
formation of the curvature radiation $l_f$.
It is important that the components
$\varepsilon_{xy}=-\varepsilon_{yx}$ and $\delta\varepsilon_{xx}$
essentially change the wave polarization. The relation between
$E_\phi$ and $E_\rho$ of the wave electric field, following from
the tensor of the dielectric permittivity (\ref{BGItensor}), is
\begin{equation}
\left(\varepsilon_{xy}+n_\rho n_\phi\right)E_\phi+\left(\delta\varepsilon_{xx}
+1-n_\phi^2\right)E_\rho=0,
\end{equation}
where $n_\rho$ and $n_\phi$ are components of the dimensionless wave vector:
${\bf n} = {\bf k}c/\omega$. For the tangent wave propagation (i.e., for $n_\rho=0$) we have
$E_\phi \simeq (\delta\varepsilon_{xx}/\varepsilon_{xy})E_\rho\simeq(c/\rho_c\tilde{\omega})E_\rho$.
As a result, a wave can produce the negative work under the electric particle
current $j_\phi$, i.e., it can be excited. It is not so if $\delta\varepsilon_{xx}=
\varepsilon_{xy}=0$ when $E_\phi=0$.
\section{Discussion}
Thus, as it was shown above, the wave polarization $[E_\rho(\rho);
\, E_\phi(\rho)]\exp\{is\phi\}$ containing one cylindrical
harmonic $s$ suggests only the Cherenkov mechanism of radiation.
In the curvature radiation mechanism of one particle in vacuum the
generated wave consists of three harmonics $s,s\pm 1$. This
property provides the exit from the phase synchronism of the wave
with the particle motion which is inherent in the bremsstrahlung
radiation. For the collective curvature radiation it is shown that
the hydrodynamical model of plasma motion along the infinite
magnetic field gives different results of the wave amplification
depending on the wave polarization. So, there is no another way to
find the polarization of exiting waves instead of calculation of
the response of the medium on an electromagnetic field, i.e., to
use the dielectric permittivity. The correct procedure of
dielectric permittivity calculation using the expansion over
cylindrical modes is also shown above. It was demonstrated that
the tensor obtained by this procedure coincides with the BGI
tensor calculated previously by another method.
In conclusion it is worth to note that unsuccessful attempts to
find the collective curvature radiation bring to the term
'curvature-drift instability' (Lyutikov et al. 1999). As was shown
the chosen simple wave polarization, i.e., one $s$-harmonic, means
only the existence of the Cherenkov mechanism of the wave
generation. In this case the centrifugal particle drift places the
significant role. Practically all curvature effects come only to
this drift. And the Cherenkov resonance on the drift motion
produces small wave amplification in better case (Kaganovich \&
Lyubarsky 2010). Stronger magnetic field produces less drift
velocity and less Cherenkov effect though the curvature of a
particle motion does not depend on the magnetic field strength at
all.
\section{Acknowledgments}
We thank A.V.~Gurevich for his interest and support.
This work was partially supported by Russian Foundation for Basic Research (Grant no.
11-02-01021).
|
\section{Introduction}\label{sec:intro}
Topological insulators have attracted significant attention in recent
years,\cite{HaKa10} especially since their experimental
realization.\cite{Koenig07} Whereas the existence of the topological
state and many of its consequences can be understood in terms of exactly
solvable, noninteracting models, the interplay of a topological band
structure and electronic correlations has become a very active field
of research. The corresponding interacting models do not have general
exact solutions, which has made computational methods one of the most
important tools. A possible experimental route to the strongly correlated
regime is based on optical lattices.\cite{PhysRevLett.107.145301}
The $Z_2$ topological band insulator (TBI), or quantum spin-Hall insulator,
closely related to the integer quantum Hall effect,\cite{HaKa10} can be realized in the
Kane-Mele (KM) model.\cite{KaMe05a,KaMe05b} The latter describes
electrons (or Dirac fermions) on the two-dimensional (2D) honeycomb lattice, with
nearest-neighbor hopping and spin-orbit coupling. Originally motivated
by graphene,\cite{KaMe05b} the spin-orbit coupling turned out to be much too
small in this material for topological effects to be observable. However, the KM model and its extension,
the Kane-Mele-Hubbard (KMH) model turn out to be a very useful theoretical
framework. In particular, the honeycomb lattice geometry provides a direct
connection to the recently discovered quantum spin liquid (QSL) phase of the
Hubbard model on the same lattice.\cite{Meng10} For the latter, the Dirac
spectrum with vanishing density of states at the Fermi level leads to a Mott
transition at a finite critical Hubbard $U$, and the QSL phase lies between a
semimetal and a magnetic insulator.\cite{Meng10} Finally, the
symmetries of the KMH model permit the application of powerful quantum Monte
Carlo (QMC) methods without a sign problem,\cite{Hohenadler10,Zh.Wu.Zh.11} so
that exact results can be obtained.
The phase diagram of the KMH model has been derived from QMC simulations,
\cite{Hohenadler10,Zh.Wu.Zh.11} and numerical results for
the extent of the QSL were presented in
Ref.~\onlinecite{Hohenadler10}. At any nonzero spin-orbit coupling, the
semimetal is replaced by the $Z_2$ TBI. In contrast, the gapped
QSL is found to be stable up to a finite critical value of the spin-orbit
interaction. Finally, the magnetic transition of the Hubbard model, between
the QSL and an antiferromagnetic Mott insulator (AFMI), is supplemented with a similar
transition between the TBI and the AFMI in a potentially different
universality class. On a qualitative level, certain aspects of the phase diagram
were obtained for example in mean-field theory,\cite{RaHu10} as well as with
cluster methods\cite{Yu.Xie.Li.11,Wu.Ra.Li.LH.11} and variational
QMC.\cite{PhysRevB.83.205122}
The understanding of the KMH model is not complete. Many of the open
questions are related to the perhaps most intriguing aspect of the model,
namely the QSL phase. The recent results from approximate cluster methods for
parameters in the QSL region of the exact phase diagram
inaccurately suggest a rather complete understanding of this exotic
phase. However, strictly speaking, any cluster method breaks translational
symmetry, so that a true QSL phase is excluded from the outset. In this
light, conclusions such as the absence of edge states, or the closing of the
single-particle gap across the transition to the TBI are not surprising,
as the QSL phase is replaced in these studies by a simple band insulator (a valence bond
crystal). The large correlation lengths (small gaps) observed
in the QSL phase in the Hubbard model\cite{Meng10}
highlight the necessity of careful interpretation of the results obtained by
cluster approximations in the context of the QSL. Interesting connections between TBIs
and QSLs are discussed in Ref.~\onlinecite{Fi.Ch.Hu.Ka.Lu.Ru.Zy.11}.
The purpose of this paper is threefold. First, we present a much more
detailed account of the QMC calculations underlying the phase diagram shown
in Ref.~\onlinecite{Hohenadler10}. Second, we extend the number of points in
parameter space and the observables calculated, in order to provide
additional insight. We also present a refined phase boundary
for the QSL phase. Third, we use the QMC method to investigate the
quantum phase transitions, especially in the light of recent theoretical
predictions.\cite{PhysRevLett.107.166806,Gr.Xu.11} We show that the TBI--AFMI
transition is in the expected 3D XY universality class, and provide evidence
for the continuous nature of the QSL--TBI and the QSL--AFMI quantum phase transitions. In
contrast to earlier work,\cite{Hohenadler10} we only consider bulk
properties. We also provide an overview of recent
work on correlation effects in topological insulators with a focus on the KMH model.
The paper is organized as follows. In Sec.~\ref{sec:model} we briefly review
the model. Details about the QMC method are presented in
Sec.~\ref{sec:method}. Section~\ref{sec:results} contains our numerical
results, beginning with the refined phase diagram, and followed by a detailed
account of the various quantum phase transitions. We end with conclusions and
an overview of open questions in Sec.~\ref{sec:conclusions}.
\section{Model}\label{sec:model}
The Hamiltonian of the KMH model can be written in the form ${H=H_\text{KM}+H_U}$, where
\begin{align}\label{eq:H}
H_\text{KM} &= -t \sum_{\las \bm{i},\bm{j} \ras}
c^{\dagger}_{\bm{i}} c^{\phantom{\dag}}_{\bm{j}} + i\,\lambda \sum_{\llas\bm{i},\bm{j}\rras}
\nu^{\phantom{\dag}}_{\bm{i}\bm{j}} c^{\dagger}_{\bm{i}} \sigma^z c^{\phantom{\dag}}_{\bm{j}} \,,\nonumber \\
H_U &= \frac{U}{2} \sum_{\bm{i}} (c^{\dagger}_{\bm{i}} c^{\phantom{\dag}}_{\bm{i}} -
1 )^2\,.
\end{align}
Here ${c^{\dagger}_{\bm{i}} = \big(c^{\dagger}_{\bm{i},\uparrow},
c^{\dagger}_{\bm{i},\downarrow}\big)}$ is a spinor of electron creation
operators, $\bm{i}$ is the position of a lattice site on the honeycomb
lattice, $\las\bm{i},\bm{j}\ras$ denotes a pair of nearest
neighbors, and $\llas \bm{i},\bm{j} \rras$ is a
pair of next-nearest-neighbor lattice sites; $\sigma^z$ is a Pauli matrix,
and $\nu_{\bm{ij}}=\pm 1$ depending on whether the hopping path defined by
the nearest-neighbor bonds connecting sites $\bm{i}$ and $\bm{j}$ bends
to the right or to the left. The complex
next-nearest-neighbor hopping term in Eq.~(\ref{eq:H}) can be related to the
spin-orbit interaction in graphene and accounts for a spin-dependent staggered
magnetic field.\cite{KaMe05b} The choice of writing the
interaction term $H_U$ in an $SU(2)$ invariant form is related to previous work on
the Hubbard model on the honeycomb lattice,\cite{Meng10} in which it was
essential to build this symmetry into the QMC method, see also
Sec.~\ref{sec:method}. In the presence of the spin-orbit term,
the $SU(2)$ spin rotation symmetry is reduced to a $U(1)$ symmetry.
Throughout this work, we use periodic boundary
conditions so that there are no edges, and take $t$ as the unit of
energy. The number of unit cells in each direction is denoted by $L$, the
total number of unit cells is $L^2$, and the total number of lattice sites is $N=2L^2$. The lattice
sizes used satisfy $L=3l$ with $l$ integer, and range from $L=3$ to $L=18$.
We exclusively consider the case of a half-filled band.
A possible Rashba term is neglected from the outset, because it would cause a
sign problem in the QMC simulations. However, a small but finite Rashba coupling
does not destroy the TBI state of the KM model.\cite{KaMe05a} The
noninteracting case $U=0$ has been solved in the original paper by Kane and
Mele.\cite{KaMe05a} Most importantly, the groundstate is a $Z_2$ TBI for any
finite spin-orbit coupling $\lambda$. The KM model is closely related to a
spinless model proposed by Haldane which shows a quantum Hall effect and
breaks time reversal invariance (TRI).\cite{Ha88} Combining two copies of the
Haldane model gives the KM model exhibiting the quantum spin-Hall effect and preserving TRI.\cite{KaMe05a,Fi.Ch.Hu.Ka.Lu.Ru.Zy.11}
Hamiltonian~(\ref{eq:H}) has been studied by means of mean-field and analytical
approaches,\cite{RaHu10,PhysRevB.82.161302,JPSJ.80.044707,We.Ka.Va.Fi.11,Va.Ma.Ho.11}
QMC simulations,\cite{Hohenadler10,Zh.Wu.Zh.11,PhysRevB.83.205122} the
variational cluster approach,\cite{Yu.Xie.Li.11} cluster dynamical mean-field
theory,\cite{Wu.Ra.Li.LH.11} and field theory.\cite{PhysRevLett.107.166806,Gr.Xu.11} A more
detailed discussion of previous results will be given in
Sec.~\ref{sec:results}.
\section{Method}\label{sec:method}
We employ a projective auxiliary-field determinant QMC algorithm similar to
Ref.~\onlinecite{Meng10}, which has previously been applied to the KMH
model.\cite{Hohenadler10,Zh.Wu.Zh.11} The method is based on the relation
\begin{equation}
\label{qmc.eq}
\langle
\Psi_0 | O |\Psi_0 \rangle = \lim_{\theta \rightarrow \infty}
\frac{
\langle
\Psi_\text{T} | e^{-\theta H/2} O e^{-\theta H/2} |
\Psi_\text{T} \rangle} {\langle \Psi_\text{T} | e^{-\theta H} |
\Psi_\text{T}
\rangle}
\end{equation}
for the expectation value of an operator $O$, with a trial wave function
$|\Psi_\text{T} \rangle $ that is required to be
nonorthogonal to the groundstate $\ket{\Psi_0}$. It is beyond the scope of this article
to describe the details of the algorithm, and the interested reader is
referred to Ref.~\onlinecite{AsEv08}. Instead, we concentrate on aspects specific
to the calculations presented here, namely the choice of the trial
wave function, the Hubbard-Stratonovich (HS) transformation,
and the absence of the minus-sign problem for a half-filled
band.
\subsection{Trial wave function}
In order to simplify the implementation, the trial wave
function is taken to be a single Slater determinant, and can hence always
be written in terms of the groundstate of a single-particle Hamiltonian
$H_\text{T}$. There are many possible choices for $\ket{\Psi_\text{T}}$. One
can for example decide to optimize the overlap with the groundstate at the
expense of symmetries.\cite{Furukawa91} Here we have preserved symmetries, and
have chosen $|\Psi_\text{T} \rangle $ to be the groundstate of the KM model,
which is defined by the first line of Eq.~(\ref{eq:H}). For $\lambda\neq 0$, the
groundstate of the noninteracting problem at half filling is
insulating. Hence, the trial wave function is nondegenerate and has all the
symmetries of the Hamiltonian. At $\lambda = 0$, the situation is more
delicate. For the considered lattice sizes, $ L = 3l$, the two nonequivalent
Dirac points are located at the Fermi surface, and the groundstate of the
noninteracting model at half filling is four-fold degenerate in each spin
sector. We lift this degeneracy by means of a twist in the boundary
condition in $H_\text{T}$ in the direction $\bm{a}_1=(1,0)$,
\begin{equation}
H_\text{T} = -t\sum_{\langle \bm{i},\bm{j} \rangle } c^{\dagger}_{\bm{i}}c^{{\phantom{\dag}}}_{\bm{j}}
\,\exp \left[\frac{ 2 \pi i }{ \Phi_0} \int_{\bm{i}}^{\bm{j}} \bm{A} \cdot d \bm{l} \right]\,,
\end{equation}
with $\bm{A} = \Phi \bm{a}_1/L $. The twist preserves translation
symmetry, so that the total momentum remains a good quantum number. In
particular, for an infinitesimal twist, the groundstate has vanishing total
momentum. Because finite values of $\Phi$ lead to a breaking of the $C_3$
lattice symmetry, the trial wave function cannot be classified according to the
irreducible representation of this group at $\lambda=0$.
After lifting possible degeneracies, $ | \Psi_\text{T} \rangle $
corresponds to the nondegenerate groundstate of $H_\text{T}$. This implies
the relation
\begin{equation}\label{eq:proj}
\lim_{\Theta \rightarrow \infty } e^{-\Theta (H_\text{T} - E_\text{T})}
= | \Psi_\text{T} \rangle \langle \Psi_\text{T} |\,,
\end{equation}
where $E_\text{T}$ is the corresponding groundstate energy.
\subsection{Hubbard-Stratonovich transformation}
We choose a HS transformation of the Hubbard term $H_U$ that couples
to the total density $n_i=n_{i\UP}+n_{i\DO}$, thereby conserving the $SU(2)$ spin symmetry for every
field configuration. In principle, HS transformations that couple to the
$z$-component of spin are also possible. However, at low temperatures, it is
often difficult to restore the $\lambda=0$ $SU(2)$ spin symmetry of the total Hamiltonian
by stochastic sampling. An $SU(2)$ spin symmetric
transformation was previously used in Refs.~\onlinecite{Assaad99,Capponi00}.
After a Trotter decomposition with imaginary time step $\Delta \tau$ to isolate
the interaction term, our HS transformation for a general operator $O$ reads
\begin{equation}\label{eq:HS}
\ensuremath e^{-\Delta\tau O^2} =
\sum_{l=\pm 1,\pm 2} \gamma(l)
\ensuremath {e}^{i\sqrt{\Delta\tau }\, \eta(l) O } + \mathcal{O}(\Delta\tau^{4}) \,,
\end{equation}
with the two functions $\gamma(l)$ and $\eta(l)$ of the auxiliary field
$l$ (with $l = \pm 1,\pm 2$) taking on the values
\begin{align}
\gamma(\pm 1) = (1+\sqrt{6}/3)/4 \,,&\quad& \eta(\pm 1) = \pm\sqrt{2\,(3-\sqrt{6})} \,,\nonumber\\
\gamma(\pm 2) = (1-\sqrt{6}/3)/4 \,,&\quad& \eta(\pm 2) = \pm\sqrt{2\,(3+\sqrt{6})} \,.
\end{align}
Equation~(\ref{eq:HS}) is an approximation to the Gaussian integral and
introduces an overall systematic error of the order
$\Delta\tau^{3}$, which is negligible in comparison to the
Trotter error of order $\Delta\tau^{2}$. The major advantage of
this approximation is that we can avoid using continuous auxiliary fields
while retaining spin rotation symmetry.
For the Hubbard interaction, we have $O = \sqrt{U/2}(n_{\uparrow}+ n_{\downarrow} -
1)$ and it is understood that the HS fields acquire space and time indices,
$l\mapsto l_{\bm{i},\tau}$.
\subsection{Absence of a sign problem}
We prove the absence of the minus-sign problem for the {\it projective} QMC
method at half filling. With the Trotter decomposition, choice of trial wave function and
HS transformation, the denominator of Eq.~(\ref{qmc.eq}) factors into spin-up
and spin-down determinants,
\begin{widetext}
\begin{equation}
\langle \Psi_\text{T} | \prod_{\tau=1}^{L_\tau} e^{-\Delta \tau
H_\text{KM} }
e^{-\Delta \tau H_U} | \Psi_\text{T} \rangle
=
\Tr \left[
\lim_{\Theta\rightarrow \infty} e^{-\Theta (H_\text{T} - E_\text{T})}
\prod_{\tau=1}^{L_\tau} e^{-\Delta \tau H_\text{KM} } e^{-\Delta \tau H_U}
\right]
=
\lim_{\Theta\rightarrow \infty}
\sum_{ \{ l_{i,\tau} \}}\prod_{\sigma}
\prod_{\tau=1}^{L_\tau} \prod_{\bm{i}} \gamma(l_{\bm{i},\tau}) W_{\sigma}\,,
\end{equation}
where we have used Eq.~(\ref{eq:proj}) to introduce a trace,
$\{l_{\bm{i},\tau}\}$ denotes an auxiliary-field configuration, and with the
weights
\begin{equation}\label{eq:Wsigma}
\begin{split}
W_{\sigma} =
\Tr \bigg[
e^{\Theta E_\text{T}}
\exp\bigg\{
-\Theta
\sum_{\bm{ij}} c^{\dagger}_{\bm{i}\si}
[h^{\phantom{\dag}}_\text{T}(\Phi)]^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}\si}
\bigg\}
\prod_{\tau=1}^{L_{\tau}}
&\exp\bigg\{
-\Delta \tau \sum_{\bm{ij}} c^{\dagger}_{\bm{i}\si}
\left( A_t + A_{\lambda,\sigma} \right)_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}\si}
\bigg\}
\\
&\times\exp\bigg\{i \sqrt{\Delta \tau U/2} \sum_{\bm{i}} \eta(l_{\bm{i},\tau})(n_{\bm{i}\si} - 1/2) \bigg\}
\bigg]\,.
\end{split}
\end{equation}
Here we introduced the notation $ H_\text{T} = \sum_{\bm{ij}\sigma}c^{\dagger}_{\bm{i}\sigma}
[h^{\phantom{\dag}}_\text{T}(\Phi)]^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}\sigma} $ and $H_\text{KM} =\sum_{\bm{ij}\sigma}c^{\dagger}_{\bm{i}\sigma}
( A_t + A_{\lambda,\sigma} )^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}\sigma} $.
Proving the absence of a negative sign problem at half
filling amounts to showing that $W_{\uparrow}^*=W_{\downarrow} $. Since the trace in
Eq.~(\ref{eq:Wsigma}) is over one spin sector, we drop the spin index
on the fermion operators to lighten the notation and obtain
\begin{align}
W^*_{\uparrow}
&=
\Tr \bigg[
e^{\Theta E_\text{T}}
\exp\bigg\{
-\Theta
\sum_{\bm{ij}} c^{\dagger}_{\bm{i}} [h^*_\text{T}(\Phi)]^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}} \bigg\}
\prod_{\tau=1}^{L_{\tau}}
\exp\bigg\{
-\Delta \tau \sum_{ij} c^{\dagger}_{\bm{i}}
( A^*_t + A^*_{\lambda,\uparrow} )^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{j}}
\bigg\}
\\\nonumber
& \qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad
\qquad\qquad \nonumber
{} \times\exp\bigg\{
-i \sqrt{\Delta \tau U/2} \sum_{\bm{i}} \eta(l_{\bm{i},\tau}) (n_{\bm{i}} -1/2)
\bigg\}
\bigg]
\\
&=\nonumber
\Tr \bigg[
e^{\Theta E_\text{T}}
\exp\bigg\{
-\Theta
\sum_{\bm{ij}} c^{\phantom{\dag}}_{\bm{i}} [h_\text{T}^*(\Phi)]^{\phantom{\dag}}_{\bm{ij}}(-1)^{\bm{i}+\bm{j}} c^{\dagger}_{\bm{j}}
\bigg\}
\prod_{\tau=1}^{L_{\tau}}
\exp\bigg\{-\Delta \tau \sum_{\bm{ij}} c^{{\phantom{\dag}}}_{\bm{i}}
( A^*_t + A^*_{\lambda,\uparrow} )^{\phantom{\dag}}_{\bm{ij}} (-1)^{\bm{i}+\bm{j}}
c^{\dagger}_{\bm{j}} \bigg\}
\\\nonumber
& \qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad
{} \times
\exp\bigg\{ i \sqrt{\Delta \tau U/2} \sum_{\bm{i}}
\eta(l_{\bm{i},\tau})( 1 - n_{\bm{i}} - 1/2) \bigg\}
\bigg]
\,.
\end{align}
\end{widetext}
The second line follows from the {\it canonical} transformation $c^{\phantom{\dag}}_{\bm{i}}
\rightarrow (-1)^{\bm{i}} c^{\dagger}_{\bm{i}} $, where the phase factor $(-1)^{\bm{i}}$
takes the value $1$ ($-1$) on sublattice A (B). The Hamiltonian $H_\text{T}$
which generates the trial wave function has nonvanishing matrix elements only
between sites on opposite sublattices. Hence $ (-1)^{\bm{i}+\bm{j}}= -1$ and
\begin{equation}\label{eq:sign1}
c^{\phantom{\dag}}_{\bm{i}} [h^*_\text{T}(\Phi)]^{\phantom{\dag}}_{\bm{ij}}(-1)^{\bm{i}+\bm{j}} c^{\dagger}_{\bm{j}}
=
c^{\dagger}_{\bm{j}} [h^*_\text{T}(\Phi)]^{\phantom{\dag}}_{\bm{ij}} c^{\phantom{\dag}}_{\bm{i}}
=
c^{\dagger}_{\bm{j}} [h_\text{T}^{\phantom{\dag}}(\Phi)]^{\phantom{\dag}}_{\bm{ji}} c^{\phantom{\dag}}_{\bm{i}}.
\end{equation}
Similarly, for the hopping term,
\begin{equation}\label{eq:sign2}
c^{\phantom{\dag}}_{\bm{i}} ( A^*_t )^{\phantom{\dag}}_{\bm{ij}}
(-1)^{\bm{i}+\bm{j}} c^{\dagger}_{\bm{j}}
=
c^{\dagger}_{\bm{j}} \left( A_t\right)^{\phantom{\dag}}_{\bm{ji}} c^{\phantom{\dag}}_{\bm{i}} .
\end{equation}
Since the spin-orbit term involves hopping between sites on the same
sublattice, we have
\begin{eqnarray}\label{eq:sign3}\nonumber
c^{\phantom{\dag}}_{\bm{i}}
( A^*_{\lambda,\uparrow} )^{\phantom{\dag}}_{\bm{ij}} (-1)^{\bm{i}+\bm{j}}
c^{\dagger}_{\bm{j}}
&=
-c^{\dagger}_{\bm{j}} \left( A_{\lambda,\uparrow} \right)^{\phantom{\dag}}_{\bm{ji}} c^{\phantom{\dag}}_{\bm{i}} \\
&=
c^{\dagger}_{\bm{j}} \left( A_{\lambda,\downarrow} \right)^{\phantom{\dag}}_{\bm{ji}} c^{\phantom{\dag}}_{\bm{i}}\,.
\end{eqnarray}
Using Eqs.~(\ref{eq:sign1})--(\ref{eq:sign3}), one sees that indeed
\begin{equation}
W^*_{\uparrow} = W_{\downarrow}\,,
\end{equation}
so that no sign problem exists at the particle-hole symmetric point of
the KMH model in the present formulation of the QMC algorithm.\cite{Hohenadler10}
The underlying reason is time reversal symmetry, which implies
$A_{\lambda,\downarrow} = - A_{\lambda,\uparrow} $.
\subsection{Measurements}
For a given auxiliary-field configuration, we have to solve a free-electron Hamiltonian
with external fields that vary in time and space. Consequently, Wick's
theorem holds, and it is sufficient to compute the single-particle Green functions
\begin{equation}
G_{\sigma}(\bm{i},\bm{j},\tau,\tau') =
- \bra{\Psi_0} T c^{\phantom{\dag}}_{\bm{i},\sigma}(\tau) c^\dag _{\bm{j},\sigma}(\tau') \ket{\Psi_0}
\end{equation}
to calculate arbitrary correlation functions. For the calculation of $G_\sigma$ we have followed
Ref.~\onlinecite{Feldbach00}. The single-particle gap $\Delta_\text{sp}$ at the Dirac
point and the (staggered) spin gap $\Delta_\text{s}$ at $\bm{q}=0$ are extracted from
fits to the corresponding Green functions.\cite{Meng10}
\subsection{Projection parameter and Trotter discretization}
The projective algorithm involves two numerical parameters, namely the
projection parameter $\theta$ and the Trotter time step $\Delta\tau$.
Both parameters were chosen such that their influence on the results is
smaller than the statistical errors. Explicitly, we used $\Delta\tau
t=0.05$ or 0.1, and $\theta t=40$--$60$.
\section{Results}\label{sec:results}
In order to better orient the discussion, we first present the phase diagram
of the KMH model in Sec.~\ref{sec:pd} together with a review of recent work,
before elaborating on the various quantum phase transitions.
\begin{figure}[t]
\includegraphics[width=0.45\textwidth]{fig_phasediagram}
\caption{\label{fig:pd} (Color online) Groundstate phase diagram of the
Kane-Mele-Hubbard model as obtained from QMC simulations. The
four phases are a $Z_2$ topological band insulator (TBI) with nonzero
single-particle (spin) gap $\Delta_\text{sp}>0$ ($\Delta_\text{s}>0$), a
semimetal (SM, $\Delta_\text{sp}=\Delta_\text{s}=0$) existing at $\lambda=0$, a quantum spin liquid
(QSL, $\Delta_\text{sp}>0$, $\Delta_\text{s}>0$), and an antiferromagnetic Mott insulator (AFMI, $\Delta_\text{sp}>0$,
$\Delta_\text{s}=0$). Magnetic order in the $z$ direction exists in the AFMI at $\lambda=0$,
but can be excluded for $\lambda/t\geq0.002$ and all values of $U/t$ shown.
Lines are quadratic fits to the QMC data points.
}
\end{figure}
\subsection{Phase diagram}\label{sec:pd}
Figure~\ref{fig:pd} shows the groundstate phase diagram of the KMH model, as
obtained from QMC simulations. In addition to the three phases of the Hubbard
model on the honeycomb lattice, the spin-orbit coupling introduces a $Z_2$ TBI. The
gapless SM phase exists only at $\lambda=0$. Whereas the TBI and the QSL
phase are fully gapped (finite single-particle gap $\Delta_\text{sp}$ and spin
gap $\Delta_\text{s}$), the magnetic phase has $\Delta_\text{sp}>0$ but
$\Delta_\text{s}=0$. Here all gaps refer to the bulk, and are not to be
confused with the metallic, gapless edge states of the TBI phase of the KMH model.
To the best of our knowledge, the QSL phase is characterized by the absence
of any local order parameter which would reflect a broken-symmetry state. It can hence be regarded as a
genuine Mott insulating state, which should be stable with respect to
small perturbations such as spin-orbit coupling. In the case of an odd number of electrons per
unit cell, the generalization of the Lieb-Schultz-Mattis theorem to two
dimensions\cite{Hasting04} suggests the presence of topological order in the
most general sense. Since
the half-filled honeycomb lattice has two electrons per unit cell, this topological
ordering still has to be numerically demonstrated or refuted. The
underlying $SU(2)\times SU(2)/Z_2$ symmetry of the Hubbard model on the
honeycomb lattice has led to the prediction of a $Z_2\times Z_2$ QSL
with mutual spin-charge statistics.\cite{PhysRevB.83.024408} Sublattice pairing states have been put
forward by various authors to account for the QSL
phase.\cite{PhysRevLett.107.087204,PhysRevB.84.024420} A canonical
consequence of the above topologically ordered phases is that, assuming
a continuous phase transition, the magnetically ordered phase would
not be a simple N\'{e}el state, thus leading to conjectures that can be tested
numerically.\cite{PhysRevLett.107.087204,PhysRevB.84.024420} Finally, in the presence of
spin-orbit coupling, the possibility of the emergence of a topological
Mott insulating phase, in which the spinons carry the topological character of
the phase, remains.\cite{PeBa10,Ru.Fi.11} For the KMH model, recent theoretical
suggestions include a $Z_2$ QSL\cite{Gr.Xu.11} and a chiral QSL.\cite{Va.Ma.Ho.11}
The boundary of the magnetic phase is obtained from the onset of long-range
antiferromagnetic order in the $xy$ plane. Longitudinal order, present at
$\lambda=0$, can be excluded in Fig.~\ref{fig:pd} for all $U/t$ and for
$\lambda/t\geq0.002$, so that the $xyz$ AFMI phase is confined to a very
small (possibly infinitesimal) interval starting at $\lambda=0$. The SM--TBI transition is evinced by the
simultaneous opening of a single-particle and a spin gap, which
as a function of $\lambda$ closely follow the $U=0$ results. The QSL--TBI transition for
intermediate Hubbard $U$ and small $\lambda$ turns out to be the most
difficult and perhaps most interesting case, with the critical values
extracted from a cusp (consistent with a closing) of the single-particle gap
$\Delta_\text{sp}$ and the spin gap $\Delta_\text{s}$. A more detailed discussion is given below.
Our numerical results suggest that the TBI phase at finite $U$ is
adiabatically connected to the TBI state of the KM model ($U=0$). Similarly, the
QSL phase is stable over a finite range of $\lambda$, in
accordance with theoretical predictions.\cite{Gr.Xu.11} Except for the
smaller range of spin-orbit couplings compared to Ref.~\onlinecite{Hohenadler10}, which is chosen here to
highlight the structure of the phase diagram around the QSL, we have obtained a number of additional points for the
phase boundary of the QSL. The refined QSL phase boundary reveals a
direct magnetic transition between the QSL and the AFMI phase at finite
$\lambda$. Our numerical data suggest the existence of a
multicritical point where the QSL, TBI and AFMI phases meet. The estimated
location of this point is $(\lambda_\text{c},U_\text{c})\approx(0.035t,4.2t)$.
Let us compare the phase diagram in Fig.~\ref{fig:pd} to other work. The
magnetic phase boundary was calculated using mean-field
theory.\cite{RaHu10} In that work, a transition from the TBI to an AFMI
phase is observed, with the critical $U$ increasing with increasing
$\lambda$ and comparable to the band width. However, the numerical values differ by up to a
factor of two. The phase diagram from unbiased QMC simulations was
presented by three of us.\cite{Hohenadler10} At that time, only one
point on the QSL--TBI phase boundary was available, and the suggested
dome-like structure of the QSL phase was based on the fact that the spin gap
takes on its maximum around $U/t=4$, in the middle of the $\lambda=0$ QSL
phase. Soon after this work, QMC results for the phase diagram
were published by Zheng \etal\cite{Zh.Wu.Zh.11} Except for the absence of
the QSL--TBI phase boundary, their phase diagram is compatible with
previous\cite{Hohenadler10} and current results
(Fig.~\ref{fig:pd}). The line $\lambda/t=0.1$ was studied by Yamaji and
Imada using variational QMC simulations,\cite{PhysRevB.83.205122} although
with rather large quantitative differences concerning the location of the
TBI--AFMI transition. The phase diagram has also been
calculated using the variational cluster approach,\cite{Yu.Xie.Li.11} and
cluster dynamical mean-field theory.\cite{Wu.Ra.Li.LH.11} Apart from the fact
that a true QSL phase is not accessible in any cluster calculation, the overall
structure of the phase diagram in these works is consistent with Fig.~\ref{fig:pd}.
The quantitative phase boundaries seem to be slightly more accurate
in the cluster dynamical mean-field case.\cite{Wu.Ra.Li.LH.11} Both papers
show a ``QSL''--TBI phase boundary whose shape is in accordance with our refined
phase diagram in Fig.~\ref{fig:pd}.
Lee\cite{PhysRevLett.107.166806} and Griset and Xu\cite{Gr.Xu.11} have
recently made predictions about the nature of some of the phase
transitions. In both works, the TBI--AFMI transition is
argued to be in the 3D XY universality class, as already hinted at in
Ref.~\onlinecite{Hohenadler10}. Griset and Xu
further suggest that both the QSL--AFMI and the QSL--TBI transitions could be
first order quantum phase transitions.\cite{Gr.Xu.11} They also highlight
the possibility of an additional, nematic order-disorder transition inside
the AFMI at $\lambda=0$, instead of a proposed chiral AF order-disorder
transition in the Hubbard model\cite{PhysRevB.84.024420} that should persist
also at $\lambda>0$.\cite{Gr.Xu.11} The phase diagram of the KMH model has
also been calculated using analytical methods.\cite{We.Ka.Va.Fi.11,Va.Ma.Ho.11}
With the number of phases and their boundaries being rather well established,
the important open questions about the phase diagram concern the nature of
the QSL and AFMI phases, and of the various phase transitions. The
structure of the phase diagram implies the existence of several distinct
quantum phase transitions: SM--TBI, TBI--AFMI, QSL--AFMI, and QSL--TBI. The
remaining SM--QSL transition only occurs at $\lambda=0$, and has been studied in
detail before.\cite{Meng10} We discuss each of these transitions below.
The remainder of this section is organized as follows. We first consider
the SM--TBI transition (at fixed $U/t=2$, see Fig.~\ref{fig:pd}) and the
TBI--AFMI transition (at fixed $\lambda/t=0.1$), for which we can
provide a fairly complete picture. From this we move on to the QSL--AFMI
transition (considering $\lambda/t=0.0125$), and finally the QSL--TBI
transition (at $U/t=4$).
\subsection{Semimetal to topological insulator transition}\label{sec:sm-tbi}
We begin with the SM--TBI transition. To this end, we keep $U/t=2$ fixed. In
the absence of interactions, the spin-orbit term breaks the sublattice
symmetry and generates a mass gap as well as a topological band structure. Due to the
underlying $U(1)$ spin symmetry, the band structure corresponds to two
Haldane models with Chern numbers of opposite sign in the two spin
sectors.\cite{KaMe05a}
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_sm-tbi}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth]{fig_sm-tbi_sp}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth]{fig_sm-tbi_spin}
\caption{\label{fig:sm-tbi} (Color online) (a) Single-particle gap $\Delta_\text{sp}$,
$2\Delta_\text{sp}$ and spin gap $\Delta_\text{s}$ as a function of $\lambda$ at $U/t=2$, across
the SM--TBI transition. The QMC results are obtained from finite-size
extrapolation using the fitting function Eq.~(\ref{eq:fit2}) at $\lambda=0$
and Eq.~(\ref{eq:fit1}) for $\lambda>0$. The numerical results shown as symbols agree
well with the corresponding gaps of the KM model ($U=0$),\cite{KaMe05b}
$\Delta_\text{sp}=3\sqrt{3}\lambda$ and $\Delta_\text{s}=2\Delta_\text{sp}$, revealing the adiabatic
connection between the TBI at $U=0$ and at $U>0$. (b) Finite-size
scaling of the single-particle gap at selected values of $\lambda/t$.
(c) Finite-size scaling of the spin gap; for $\lambda>0$, we neglect
the $L=3$ results in the extrapolation.}
\end{figure}
Figure~\ref{fig:sm-tbi}(a) shows QMC results for the
single-particle gap $\Delta_\text{sp}$ and the spin gap $\Delta_\text{s}$ as a function of
$\lambda$ at $U/t=2$. Starting in the SM
phase at $\lambda=0$, where both gaps are zero, $\Delta_\text{sp}$ and $\Delta_\text{s}$ become nonzero for any finite
$\lambda$, and increase with increasing spin-orbit coupling. The QMC results
at $U/t=2$ closely follow the corresponding gaps
for the noninteracting case $U=0$,\cite{KaMe05b} $\Delta_\text{sp}=3\sqrt{3}\lambda$ and
$\Delta_\text{s}=2\Delta_\text{sp}$. Interaction effects manifest themselves as a minor suppression
of both gaps compared to their noninteracting values, especially at
larger $\lambda/t$, and by the spin gap falling below $2\Delta_\text{sp}$.
From these results, we draw the following conclusions. First, the SM phase of
the Hubbard model is unstable at finite $\lambda$, and hence only exists for
$\lambda=0$, as indicated in Fig.~\ref{fig:pd}. Second, the very small
deviations in the dependence of the gaps on $\lambda$ compared to the
noninteracting case suggest that the TBI phase at $U>0$ is essentially the
same as at $U=0$, provided $U$ remains small enough to avoid the magnetic
transition. This finding suggests that the two states are adiabatically
connected. The minor role of bulk interactions inside the TBI phase may
be regarded as a consequence of the single-particle energy gap,\cite{HaKa10}
and has been exploited to develop an effective model of the helical edges
with a Hubbard $U$ only at the edge sites of a ribbon.\cite{Hohenadler10,Ho.As.11}
The results for $\Delta_\text{sp}$ and $\Delta_\text{s}$ in Fig.~\ref{fig:sm-tbi}(a) are obtained
from finite-size scaling, as shown for selected values of $\lambda/t$ in
Figs.~\ref{fig:sm-tbi}(b) and (c). Whereas $\Delta_\text{sp}$ shows the familiar
monotonic decrease with increasing system size, the spin gap reveals an
unusual finite-size scaling behavior. Deep in the TBI phase [\eg, the top
curve in Fig.~\ref{fig:sm-tbi}(c) corresponding to $\lambda/t=0.03$],
$\Delta_\text{s}$ systematically increases with increasing system size $L$. For small
$\lambda$ (for example, $\lambda/t=0.0125$) and small $L$, the scaling
behavior is more complex, and only the system sizes beyond the crossover have
been used in the extrapolation. The increase of $\Delta_\text{s}$ with increasing $L$
is a correlation effect; the spin gap is independent of $L$ for $U=0$.
The observed increase of $\Delta_\text{s}$ with system size can be reproduced using first-order
perturbation theory in $U$.
A possible physical explanation is inspired by the observation that for small $L$,
$\Delta_\text{s}(L)< 2\Delta_\text{sp}(L)$, see Figs.~\ref{fig:sm-tbi}(b) and (c). In contrast, the
extrapolated values almost match the relation for the noninteracting case,
$\Delta_\text{s}\approx 2\Delta_\text{sp}$, as shown in Fig.~\ref{fig:sm-tbi}(a). The strongly
suppressed spin gap for small system sizes indicates pronounced particle-hole
binding, driven by correlation-induced magnetic fluctuations. If the length
scale of these fluctuations, which are a precursor of the magnetic transition
at $U_\text{c}$, exceeds the system size, the spin gap is expected to be suppressed
similar to the magnetic phase where $\Delta_\text{s}\to0$ as $L\to\infty$. Increasing $L$
beyond the correlation length will restore the behavior expected for the
weakly or noninteracting TBI phase. We will see below
(Fig.~\ref{fig:qsl-tbi-sp}) that for larger $U/t=4$, the spin gap is
suppressed to values much below $2\Delta_\text{sp}$ even in the thermodynamic limit.
Although a complete understanding of this effect is currently missing,
we regard the unusual spin gap scaling as a signature of a correlated TBI.
\subsection{Topological insulator to antiferromagnet transition}\label{sec:tbi-afmi}
At large $U/t$, the TBI phase of the half-filled KMH model undergoes a
transition to an AFMI, and TRI is spontaneously broken.\cite{RaHu10}
In the strong-coupling limit, $U / t \gg 1$, the charge degrees of freedom are frozen
and one can derive an effective spin model with antiferromagnetic nearest-neighbor Heisenberg
exchange $J=4t^2/U$ that promotes isotropic magnetic order in the $xy$ and
$z$ directions. The spin-orbit term of the KMH model reduces the $SU(2)$ spin symmetry of the
Hubbard model to a $U(1)$ symmetry corresponding to conservation of the total
$z$-component of spin. Second-order perturbation theory gives an exchange
interaction $J'=4\lambda^2/U$ between next-nearest
neighbors. Importantly, the exchange is antiferromagnetic in the longitudinal
direction, $J' S^z_{\bm{i}} S^z_{\bm{j}}$, but ferromagnetic in the
transverse direction, $-J'(S^x_{\bm{i}} S^x_{\bm{j}}+S^y_{\bm{i}} S^y_{\bm{j}})$.\cite{RaHu10}
Combining all the exchange terms, magnetic order in the $z$ direction becomes
frustrated, and the system favors an easy-plane N\'eel state. The so-called
KM-Heisenberg model was recently studied analytically.\cite{Va.Ma.Ho.11}
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_safxy}
\caption{ \label{fig:tbi-afmi} (Color online)
Finite-size scaling of the rescaled magnetic structure factor
$S_\text{AF}^{xy}/N$ defined in Eq.~(\ref{eq:SAF}) at
$\lambda/t=0.1$ for different values of $U/t$, across the TBI--AFMI
transition. The curves (representing polynomial fits) extrapolate to zero for
$U/t<4.9$, and to a finite value for $U/t\geq5.0$, giving the critical value $U_\text{c}/t=4.95(5)$.
A more accurate estimate $U_\text{c}/t=4.96(4)$ is obtained from
Fig.~\ref{fig:3DXY_SAFxy}(a). The inset shows the order parameter $m^{xy}$ as obtained from
extrapolation to the thermodynamic limit.}
\end{figure}
From the above considerations, $xy$ order is expected both for $\lambda=0$
and $\lambda\neq0$. Hence, the phase boundary of the AFMI phase can be
determined from the onset of transverse long-range magnetic order by
monitoring the transverse structure factor
\begin{align}\label{eq:SAF}
S_\text{AF}^{xy}
&\equiv \sum_{\alpha} [S_\text{AF}^{xy}]^{\alpha\alpha}\,,
\\\nonumber
[S_\text{AF}^{xy}]^{\alpha\beta}
&= \frac{1}{L^2}\sum_{\bm{r}\bm{r}'} (-1)^{\alpha} (-1)^{\beta}
\langle \Psi_0
| S^{+}_{\bm{r}\alpha} S^{-}_{\bm{r}'\beta} + S^{-}_{\bm{r}\alpha} S^{+}_{\bm{r'}\beta} | \Psi_0 \rangle\,.
\end{align}
Here $\bm{r},\bm{r}'$ denote unit cells, $\alpha,\beta\in\{A,B\}$ are sublattice indices,
$(-1)^\alpha=1$ ($-1$) for $\alpha=A$ ($B$), and we have taken the trace of
the corresponding $2\times2$ matrix of the structure factor.
The quantity $S_\text{AF}^{xy} /N$ (with $N=2L^2$) extrapolates to zero below
$U_\text{c}(\lambda)$, but takes on a finite value in the thermodynamic
limit for $U\geqU_\text{c}(\lambda)$, which is inside the AFMI phase of
Fig.~\ref{fig:pd}.\cite{Hohenadler10,Zh.Wu.Zh.11} It is also related
to the transverse magnetization via $m_{xy}^2=S_\text{AF}^{xy} /N$.
Numerical results for $\lambda/t=0.1$ are shown in Fig.~\ref{fig:tbi-afmi};
the transition is most obvious from the extrapolated order
parameter shown in the inset. The extrapolation of
$S_\text{AF}^{xy} /N$ in system size gives a critical value $U_\text{c}/t=4.95(5)$. This value
agrees with the slightly more accurate estimate $U_\text{c}/t=4.96(4)$ which follows
from the intersect of curves for different system sizes in Fig.~\ref{fig:3DXY_SAFxy}(a).
However, this scaling analysis (see below for more details) relies on the knowledge
of the universality class of the transition. By performing calculations at
different $\lambda/t$, we can determine the magnetic phase boundary, and we
find good agreement with previous exact
simulations at $\lambda=0$\cite{Meng10} and
$\lambda>0$.\cite{Hohenadler10,Zh.Wu.Zh.11} Our QMC results (not
shown) further exclude the presence of longitudinal magnetic order along the
entire TBI--AFMI phase boundary in Fig.~\ref{fig:pd} and up to $U/t=8$.
The TBI--AFMI transition is also reflected in the single-particle gap.
Because the onset of long-range magnetic order at the TBI--AFMI transition
spontaneously breaks TRI, the transition from the TBI to the
nonadiabatically connected AFMI can in principle occur without
closing any excitation gaps. Instead, the transition manifests itself in
$\Delta_\text{sp}$ as a cusp at $U_\text{c}/t=4.95(5)$, visible in Fig.~\ref{fig:tbi-afm-gaps}(a). The
results are for the same value of $\lambda/t=0.1$ considered in Fig.~\ref{fig:tbi-afmi}.
A similar signature can be reproduced already on the mean-field level, although
with only a kink instead of a cusp at the critical point. Results for the gap and the
mean-field order parameter are presented in the inset of
Fig.~\ref{fig:tbi-afm-gaps}(a). Figure~\ref{fig:tbi-afm-gaps}(a) also shows the
closing of the spin gap $\Delta_\text{s}$ at $U_\text{c}$; the results
were obtained from the finite-size scaling shown in the inset of Fig.~\ref{fig:3DXY_gap}(a).
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_gaps}\vspace*{0.75em}
\includegraphics[width=0.425\textwidth]{fig_dfdu_lambda0.1.eps}
\caption{\label{fig:tbi-afm-gaps} (Color online) (a) Single-particle gap
$\Delta_\text{sp}$ and spin gap $\Delta_\text{s}$ as a function of $U$ at
$\lambda/t=0.1$. The values shown were obtained from an extrapolation to
the thermodynamic limit. The dip in $\Delta_\text{sp}$ and the closing of
$\Delta_\text{s}$ are consistent with $U_\text{c}/t=4.95(5)$. The inset shows
the mean-field results for the single-particle gap and the magnetic
order parameter. (b) Energy derivative with respect to $U$
[Eq.~(\ref{eq:dfdU})] across the TBI--AFMI transition at
$\lambda/t=0.1$ [$U_\text{c}/t=4.95(5)$]. }
\end{figure}
In Fig.~\ref{fig:tbi-afm-gaps}(b), we present numerical data for the
energy derivative
\begin{equation}\label{eq:dfdU}
\frac{\partial F}{\partial U}
=
\las
\mbox{$\frac{1}{2}$}
\sum_{\bm{i}} (c^{\dagger}_{\bm{i}} c^{\phantom{\dag}}_{\bm{i}} - 1 )^2
\ras \,,
\end{equation}
corresponding to the expectation value of the interaction term or, equivalently,
the average double occupation, at $\lambda/t=0.1$. The continuous variation
of this quantity across $U_\text{c}/t=4.95(5)$ suggests a continuous transition.
Having established the phase boundary of the magnetic transition at large $U/t$,
we now consider the universality class. Given the remaining $U(1)$ spin
symmetry in the presence of spin-orbit coupling, the transition is expected to be in the 3D XY
universality class. An intuitive picture is based on local magnetic
moments, which already exist in the magnetically disordered phase for
$U>0$, and order at $U_\text{c}$. The onset of phase coherence at $U=U_\text{c}$
corresponds to a $U(1)$ symmetry breaking. This scenario
is in accordance with the behavior of the spin gap $\Delta_\text{s}$ in
Fig.~\ref{fig:tbi-afm-gaps}. The excitons are massive in the disordered phase
($U<U_\text{c}$), but condense in the ordered phase ($U\geqU_\text{c}$) where
$\Delta_\text{s}=0$.
The conjectured 3D XY universality can be tested using the zero-temperature,
finite-size scaling forms
\begin{equation}\label{eq:scaling:S}
S_\text{AF}^{xy}/N = L^{-2\beta/\nu} f_1[(U-U_\text{c}) L ^{1/\nu}]
\end{equation}
and
\begin{equation}\label{eq:scaling:D}
\Delta_\text{s}/t = L^{-z} f_2[(U-U_\text{c}) L ^{1/\nu}]\,.
\end{equation}
Here $f_1$ and $f_2$ are dimensionless functions. The relevant
critical exponents for the 3D XY model are
$z=1$, $\nu=0.6717(1)$ and $\beta=0.3486(1)$.\cite{PhysRevB.74.144506}
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_safxy_intersect}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_safxy_collapse}
\caption{\label{fig:3DXY_SAFxy} (Color online)
Rescaled transverse magnetic structure factor $S_\text{AF}^{xy}/N$
defined in Eq.~(\ref{eq:SAF}) as a function of $U$ at $\lambda/t=0.1$,
for different lattice sizes $L$. Assuming the scaling
form~(\ref{eq:scaling:S}), (a) shows $L^{2\beta/\nu}S_\text{AF}^{xy}/N$. The intersection of
curves for different system sizes yields $U_\text{c}/t=4.96(4)$ for the critical point.
(b) The scaling collapse obtained by plotting $L^{2\beta/\nu}
S_\text{AF}^{xy}/N$ as a function of $L^{1/\nu}
(U-U_\text{c})/U_\text{c}$. The QMC data are fully consistent with the critical
exponents $z=1$, $\nu=0.6717(1)$ and $\beta=0.3486(1)$ of the 3D XY
model.\cite{PhysRevB.74.144506} }
\end{figure}
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_ds_intersect}\vspace*{0.45em}
\includegraphics[width=0.425\textwidth]{fig_tbi-afmi_ds_collapse}
\caption{\label{fig:3DXY_gap} (Color online)
Spin gap $\Delta_\text{s}$ as a function of $U$ at
$\lambda/t=0.1$, for different lattice sizes $L$. Given the scaling
form Eq.~(\ref{eq:scaling:D}), (a) shows $L^{z} \Delta_\text{s}$. The intersection
of curves for different $L$ gives $U_\text{c}/t=4.96(4)$, consistent
with Fig.~\ref{fig:3DXY_SAFxy}(a). The inset shows the finite-size
scaling of $\Delta_\text{s}$. (b) Scaling collapse obtained by plotting $L^{z}\Delta_\text{s}$ as a
function of $L^{1/\nu}(U-U_\text{c})/U_\text{c}$. The QMC data are consistent with the 3D
XY exponents $z=1$, $\nu=0.6717(1)$ and
$\beta=0.3486(1)$.\cite{PhysRevB.74.144506}
}
\end{figure}
Using the same value $\lambda/t=0.1$ as before, we show in
Fig.~\ref{fig:3DXY_SAFxy}(a) $L^{2\beta/\nu} S_\text{AF}^{xy}/N$ as a function of $U$ for different system
sizes $L$. If the scaling form Eq.~(\ref{eq:scaling:S}) with the critical
exponents of the 3D XY model is correct, we expect to see an intersect
of curves for different $L$ at $U=U_\text{c}$. As shown in
Fig.~\ref{fig:3DXY_SAFxy}(a), this prediction is indeed borne out by the QMC data,
and we deduce $U_\text{c}/t=4.96(4)$, in agreement with Fig.~\ref{fig:tbi-afmi}. Replotting
$L^{2\beta/\nu} S_\text{AF}^{xy}/N$ as a function of $L^{1/\nu}
(U-U_\text{c})/U_\text{c}$ in Fig.~\ref{fig:3DXY_SAFxy}(b)
produces a clean scaling collapse onto a single
curve. Figure~\ref{fig:3DXY_SAFxy} hence demonstrates that the assumption of
3D XY behavior is fully consistent with the QMC data.
Figure~\ref{fig:3DXY_gap} shows a similar analysis for the spin gap $\Delta_\text{s}$,
using the scaling form~(\ref{eq:scaling:D}). Although the statistical quality
of the data is not quite as good as for the structure factor, we again find
satisfactory scaling (in particular, there is no noticeable drift of the
intersect with increasing $L$) and the same $U_\text{c}$ using the 3D XY critical exponents.
Based on the existence of a $U(1)$ spin symmetry throughout the TBI phase,
we expect the 3D XY behavior found at $\lambda/t=0.1$ to be generic for this transition,
in agreement with previous predictions.\cite{Hohenadler10,Gr.Xu.11,PhysRevLett.107.166806}
The finite-size corrections to the 3D XY scaling
behavior become more pronounced on approaching the possible multicritical
point, as verified explicitly for $\lambda/t=0.05$. According to Griset and
Xu,\cite{Gr.Xu.11} the observed 3D XY behavior at $\lambda>0$ suggests the
absence of a chiral AF order-disorder transition inside the AFMI phase even for
$\lambda=0$.\cite{PhysRevB.84.024420}
\subsection{Spin liquid to antiferromagnet transition}\label{sec:qsl-afmi}
The refined phase boundary of the QSL phase shown in Fig.~\ref{fig:pd}
establishes the existence of a QSL--AFMI transition at finite $\lambda$,
in addition to the $\lambda=0$ transition studied before.\cite{Meng10}
Since the present work is concerned with the KMH model, we only consider
finite values $\lambda>0$ here.
The simple picture of the magnetic transition as an ordering transition
of magnetic moments (or exciton condensation) discussed in the context
of the TBI--AFMI transition cannot straightforwardly be applied to
the QSL--AFMI transition. For example, a $Z_2$ spin liquid exhibits
charge fractionalization, and therefore has no well-defined magnetic
modes. Fractionalization could lead to an unusually large anomalous
dimension.\cite{Is.Me.Ha.11} On the other hand, if the QSL phase was adiabatically
connected to a simple band insulator (without charge fractionalization), the
transition is again expected to be of the 3D XY type, similar to the TBI--AFMI transition.
\begin{figure}[t]
\includegraphics[width=0.425\textwidth,clip]{fig_qsl-afmi_xy}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth,clip]{fig_qsl-afmi_zz}
\caption{ \label{fig:xyorder} (Color online)
(a) Finite-size scaling of the rescaled magnetic structure factor
$S_\text{AF}^{xy}/N$ defined in Eq.~(\ref{eq:SAF}) at
$\lambda/t=0.0125$ for different values of $U/t$, across the QSL--AFMI
transition. The data suggest a magnetic transition at
$U_\text{c}/t=4.3(2)$. Lines correspond to polynomial fits.
(b) Same as in (a) but showing the longitudinal structure factor $S_\text{AF}^{zz}$ defined in
Eq.~(\ref{eq:SAFzz}).
In contrast to (a), there is no long-range order over the
range of $U/t$ values considered.
}
\end{figure}
Due to the small size of the spin gap in the QSL phase, the
extrapolation of the order parameter~(\ref{eq:SAF}) to the thermodynamic
limit is much more delicate than for the TBI-AFMI transition. In particular,
a scaling analysis along the lines of Figs.~\ref{fig:3DXY_SAFxy} and~\ref{fig:3DXY_gap}
is not conclusive with the currently available system sizes.
We first address the question of longitudinal magnetic order.
The phase diagram presented by Yu \etal,\cite{Yu.Xie.Li.11} based on results
from the variational cluster approach, shows an extended region inside the AFMI phase
in which the authors claim that magnetic order exists both in the $xy$ plane and
in the $z$ direction. For $\lambda=0$, this region is argued to extend all the
way to the QSL--AFMI phase boundary, leading to a simultaneous onset of
transverse and longitudinal order at $U_\text{c}$. At $\lambda>0$, Yu
\etal\cite{Yu.Xie.Li.11} find a transition from the TBI to an $xy$ ordered
AFMI at $U_\text{c}$, and an onset of $z$ order at even larger values
of $U$. Hence, for $\lambda>0$, there would be an additional crossover (no symmetry breaking)
inside the AFMI phase. Whereas $z$ order is known to exist in the Hubbard
model ($\lambda=0$), this result is surprising in the light of the strong-coupling
picture mentioned above, in which antiferromagnetic correlations in the $z$
direction are frustrated by the interplay of hopping $t$ and spin-orbit coupling
$\lambda$.
To clarify the situation, we use unbiased QMC simulations and calculate the
transverse structure factor [Eq.~(\ref{eq:SAF})] as well as the longitudinal structure factor
\begin{align}\label{eq:SAFzz}
S_\text{AF}^{zz}
&\equiv \sum_\alpha [S_\text{AF}^{zz}]^{\alpha\alpha}\,,
\\\nonumber
[S_\text{AF}^{zz}]^{\alpha\beta}
&= \frac{1}{L^2}\sum_{\bm{r}\bm{r}'} (-1)^{\alpha} (-1)^{\beta}
\langle \Psi_0
| S^{z}_{\bm{r}\alpha} S^{z}_{\bm{r}'\beta} | \Psi_0 \rangle\,,
\end{align}
at $\lambda/t=0.0125$. The results are shown in Fig.~\ref{fig:xyorder}. The
onset of transverse magnetic order is visible from the finite-size extrapolation of
$S_\text{AF}^{xy}/N$ depicted in Fig.~\ref{fig:xyorder}(a),
and the critical value $U_\text{c}/t=4.3(2)$ is shown in the phase diagram in
Fig.~\ref{fig:pd}. However, as revealed by Fig.~\ref{fig:xyorder}(b),
there is no long-range order in the longitudinal direction even for large
values of $U/t=8$. We have carried out simulations down to $\lambda/t=0.002$,
where longitudinal order would be most favorable, but found no $z$ order for the
$U$ range shown in Fig.~\ref{fig:pd}. Hence, an extended region of
$z$ order as suggested by Ref.~\onlinecite{Yu.Xie.Li.11} does not exist, and
the phase diagram is instead given by Fig.~\ref{fig:pd}, with a very narrow,
possibly infinitesimal, region of coexisting longitudinal and transverse order near $\lambda=0$. The
discrepancy between our exact numerical results and those of the variational cluster approach is most
likely a consequence of the very small cluster sizes used for the latter.
Although the strong-coupling picture with exchange constants $J,J'$
is not justified for intermediate $U$, the frustration in the $z$ direction
qualitatively explains the absence of longitudinal order found
numerically. The purely in-plane magnetic order agrees with
field-theory predictions for the KMH model.\cite{Gr.Xu.11}
Using field theory arguments, Griset and Xu\cite{Gr.Xu.11} suggested
the possibility that the QSL--AFMI transition could be first order.
To test this hypothesis, we show in Fig.~\ref{fig:qsl-tbi-docc}(a)
the energy derivative $ \partial F/\partial U$ [Eq.~(\ref{eq:dfdU})].
We do not find any sign of discontinuous behavior near $U_\text{c}$, which suggests
that the transition is continuous. However, we cannot exclude
the possibility of a weakly first-order transition. We have also
calculated $ \partial F/\partial U$ at $\lambda/t=0.04$ (close to the
multicritical point) and found no signature of discontinuous behavior, see Fig.~\ref{fig:qsl-tbi-docc}(b).
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_qsl-afmi_dfdu_lambda0.0125.eps}\vspace*{0.75em}
\includegraphics[width=0.425\textwidth]{fig_qsl-afmi_dfdu_lambda0.04.eps}
\caption{\label{fig:qsl-tbi-docc} (Color online) Energy derivative
with respect to $U$ [Eq.~(\ref{eq:dfdU})] across the QSL--AFMI transition at
$\lambda/t=0.0125$ [$U_\text{c}/t=4.3(2)$], and at $\lambda/t=0.04$
[$U_\text{c}/t=4.3(2)$], close to the multicritical point. There are no
signs of a first-order transition.}
\end{figure}
\begin{figure}[ht]
\includegraphics[width=0.425\textwidth,clip]{fig_qsl-tbi_sp}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth,clip]{fig_qsl-tbi_spin}\vspace*{0.5em}
\includegraphics[width=0.425\textwidth,clip]{fig_qsl-tbi_spin_scaling}
\caption{\label{fig:qsl-tbi-sp} (Color online) (a) Single-particle gap
$\Delta_\text{sp}$ and (b) spin gap $\Delta_\text{s}$ and inverse spin correlation length
$1/\xi_\text{s}'$ as a function of $\lambda$ at $U/t=4$, across the QSL--TBI
transition. $\Delta_\text{sp}$ is obtained from finite-size scaling using
Eq.~(\ref{eq:fit1}) for $\lambda/t\geq0.045$ and Eq.~(\ref{eq:fit2})
for $\lambda/t<0.045$. The cusp defines the critical coupling $\lambda_\text{c}/t=0.030(1)$.
The inset in (a) shows the scaling for selected values of $\lambda/t$.
$\Delta_\text{s}$ is obtained from finite-size scaling using Eq.~(\ref{eq:fit1}), see
(c). $\xi_\text{s}'$ is extracted from fits to the spin-spin correlation
function defined in Eq.~(\ref{eq:spinspin}) at $r=L/2$,
see inset in (b).}
\end{figure}
\subsection{Spin liquid to topological insulator transition}\label{sec:qsl-tbi}
A characteristic feature of both the QSL and the TBI phase is the absence of
broken symmetries. Therefore, the QSL--TBI transition cannot be tracked by
a local order parameter. In previous work,\cite{Hohenadler10} the
critical point $\lambda_\text{c}$ at $U/t=4$ was determined from the behavior of the
single-particle gap. Here we discuss the underlying procedure in detail,
present new results with improved resolution of the critical point and based
on larger system sizes up to $L=18$, and refine the phase boundary by determining
the critical point at two other values of $U/t$. Moreover, we show numerical
results for the spin gap, and address the possibility of a first-order transition.
Figure~\ref{fig:qsl-tbi-sp}(a) shows the single-particle gap $\Delta_\text{sp}$ as
a function of $\lambda$ at $U/t=4$. The data points are obtained from
extrapolation to the thermodynamic limit, as illustrated
in the inset. Deep in the TBI phase ($\lambda>\lambda_\text{c}$), we use the fitting function
($\alpha=\text{sp}$, $\text{s}$)
\begin{equation}\label{eq:fit1}
\Delta_\alpha(L)/t
=
a + e^{-L/\xi_\alpha} (b/L + c/L^2 )\,,
\end{equation}
with a correlation length $\xi_{\alpha}$. On approaching $\lambda_\text{c}$ from
above, the correlation length set by the single-particle gap, $\xi_\text{sp}$,
increases and exceeds $L/2=9$ ($L=18$ being our largest system size) for
$\lambda/t\approx0.04$. For $\lambda/t$ smaller than 0.04, we use
\begin{equation}\label{eq:fit2}
\Delta_{\alpha}(L)/t = a+ b/L + c/L^2 \,.
\end{equation}
As a function of $\lambda$, the extrapolated single-particle gap in
Fig.~\ref{fig:qsl-tbi-sp}(a) initially
decreases when starting from the QSL at $\lambda=0$, reveals
a cusp centered at $\lambda_\text{c}/t=0.030(1)$, and increases
rather quickly with increasing $\lambda$ for $\lambda>\lambda_\text{c}$. As in previous
work,\cite{Hohenadler10} we take the location of the cusp to define the
critical point $\lambda_\text{c}$ of the QSL--TBI transition. We will argue below that
the data are consistent with a closing of the gap at $\lambda_\text{c}$, and that the
cusp is a result of finite-size effects. For $\lambda/t\geq0.045$, the
larger system sizes now available result in larger values of $\Delta_\text{sp}$ compared
to previous work.\cite{Hohenadler10}
A similar analysis can be carried out for the spin gap $\Delta_\text{s}$ using
Eq.~(\ref{eq:fit1}) for $\lambda>\lambda_\text{c}$. Similar to the SM--TBI transition
discussed above, the spin gap shows an unusual finite-size scaling inside the
TBI phase, as shown in Fig.~\ref{fig:qsl-tbi-sp}(c). The system sizes
required to see saturation (\ie, the magnetic correlation lengths) are
significantly larger at $U/t=4$ than at $U/t=2$,
cf. Fig.~\ref{fig:sm-tbi}(c). The extrapolated values of $\Delta_\text{s}$ for
$\lambda>\lambda_\text{c}$ are shown in Fig.~\ref{fig:qsl-tbi-sp}(b). Comparing
$\Delta_\text{sp}$ and $\Delta_\text{s}$ [Figs.~\ref{fig:qsl-tbi-sp}(a) and \ref{fig:qsl-tbi-sp}(b)],
we see that in contrast to $U/t=2$ [Fig.~\ref{fig:sm-tbi}(a)] we have
$\Delta_\text{s}<2\Delta_\text{sp}$ in the TBI phase at $U/t=4$. The suppressed spin gap indicates
substantial particle-hole binding. In the QSL phase, the small values
of the spin gap make an accurate determination very challenging; $\Delta_\text{s}$ is
largest at $\lambda=0$, where it was previously determined as $\Delta_\text{s}/t=0.023(5)$.\cite{Meng10}
Figure~\ref{fig:qsl-tbi-sp}(b) reveals that the behavior of the spin gap for
$\lambda>\lambda_\text{c}$ is very similar to that of $\Delta_\text{sp}$. As a consistency check, we
also show the inverse spin correlation length $\xi_\text{s}$. Assuming the
form
\begin{equation}\label{eq:spinspin}
S^{xx}(r) = \langle S^x_{\bm{r}} S^x_{\bm{0}} \rangle = e^{-r/\xi_\text{s}'} ( a/r + b/r^2)
\end{equation}
for the real-space transverse spin-spin correlation function, and taking the largest available
distance $r=L/2$ for each system size [see inset of
Fig.~\ref{fig:qsl-tbi-sp}(b)], the dependence of $1/\xi_\text{s}'$ on
$\lambda$ is in good agreement with $\Delta_\text{s}$ and $\Delta_\text{sp}$. We
only show the values $\xi_\text{s}'\leq L/2$.
For a noninteracting $Z_2$ TBI, there is a simple relation between the
excitation gaps in the single-particle sector and, \eg, in the spin and
particle-hole channels. In the presence of (strong) interactions, these
relations may be modified, and we indeed find $\Delta_\text{s}<2\Delta_\text{sp}$ in the TBI phase
above the QSL--TBI transition, as well as in the QSL phase at
$\lambda=0$.\cite{Meng10} The argument that $\Delta_\text{sp}$ has to close across a
transition that involves a change of the topological index\cite{HaKa10} holds
only for the noninteracting case. In general, it is not clear which
excitation gaps (one or more) close if the states on either side of the
transition are not adiabatically connected. For example, in the interacting
Haldane model,\cite{Va.Su.Ri.Ga.11} there is an exact degeneracy of the three
lowest states at the TBI to charge density wave transition. As a result, the
first and second excitation gaps ($E_1-E_0$ and $E_2-E_0$) close, but the
single-particle gap $\Delta_\text{sp}$ shows only a cusp at the critical point.
As argued in previous work,\cite{Hohenadler10} the results for $\Delta_\text{sp}$ are
consistent with a vanishing of the single-particle gap at $\lambda_\text{c}$. Furthermore, the results in
Fig.~\ref{fig:qsl-tbi-sp} reveal that $\Delta_\text{sp}$ and $\Delta_\text{s}$ behave very similar on
approaching $\lambda_\text{c}$ from above, and we may therefore expect to see a
simultaneous closing of $\Delta_\text{sp}$ and $\Delta_\text{s}$. Such a gap closing suggests
different Chern numbers for the TBI and QSL phases. Additionally,
the quick, almost linear opening of the gaps
for $\lambda>\lambda_\text{c}$ is reminiscent of Fig.~\ref{fig:sm-tbi}(a) for the SM--TBI
transition, suggesting that a non-TBI phase (the QSL) exists at small values of
$\lambda$, and that a transition to the TBI phase takes place at $\lambda_\text{c}>0$.
This picture confirms the expectation that the fully gapped QSL phase should
be stable under a small perturbation in the form of the spin-orbit term.
We attribute the small but nonzero values of the gaps at $\lambda_\text{c}$
to finite-size effects. Although we used the same range of
system sizes (up to $L=18$) as for $\lambda=0$,\cite{Meng10} the larger
correlation lengths in the present case, especially in the spin channel, make
the analysis significantly harder. On approaching $\lambda_\text{c}$ from above, the
correlation lengths exceed the largest distance available on the clusters
used. If we consider only the data points for which the correlation lengths
fit on the largest system, \ie the range $\lambda/t\geq0.045$ in
Figs.~\ref{fig:qsl-tbi-sp}(a) and \ref{fig:qsl-tbi-sp}(b),
the functional form of $\Delta_\text{sp}$, $\Delta_\text{s}$ and $1/\xi_\text{s}'$ strongly suggests
a closing of the gaps very close to $\lambda_\text{c}/t=0.030(1)$. The fact that the finite-size
scaled gaps in Figs.~\ref{fig:qsl-tbi-sp}(a) and \ref{fig:qsl-tbi-sp}(b) saturate at a finite
value close to $\lambda_\text{c}$ is therefore likely to be a result of insufficiently large
system sizes. The latter require that we switch to the polynomial fitting
function~(\ref{eq:fit2}) close to $\lambda_\text{c}$ for $\Delta_\text{sp}$, and do not permit
a reliable calculation of $\Delta_\text{s}$ or $\xi_\text{s}'$ close to $\lambda_\text{c}$. The question if the gaps
close or not cannot be answered using approximate cluster
calculations,\cite{Yu.Xie.Li.11,Wu.Ra.Li.LH.11} because such methods are not
capable of describing a true QSL phase.
To determine the shape of the QSL phase boundary, we have calculated the
single-particle gap for two other values of $U/t$. The critical
values $\lambda_\text{c}$ are again defined by the location of the cusp in $\Delta_\text{sp}$.
We find $\lambda_\text{c}/t=0.025(2)$ for $U/t=3.8$ and $\lambda_\text{c}/t=0.032(2)$
for $U/t=4.1$.
Recent theoretical work based on a $1/N$ expansion predicts the possibility
of a first-order QSL--TBI transition.\cite{Gr.Xu.11} To test this prediction,
we show in Fig.~\ref{fig:qsl-tbi-ener} the quantity
\begin{equation}\label{eq:dfdlambda}
\frac{\partial F}{\partial \lambda}
=
\las\,
i \! {\ensuremath \sum_{\llas\bm{i},\bm{j}\rras}}
\nu^{\phantom{\dag}}_{\bm{i}\bm{j}}
c^{\dagger}_{\bm{i}}
{\sigma}^z c^{\phantom{\dag}}_{\bm{j}}
\ras
\,,
\end{equation}
corresponding to the expectation value of the spin-orbit term in Eq.~(\ref{eq:H}).
For the range of system sizes, and on the very fine grid of $\lambda$ values,
there is no sign of a discontinuity. Again, we cannot rule out the possibility
of a weakly first-order transition.
\begin{figure}[t]
\includegraphics[width=0.425\textwidth]{fig_qsl-tbi_dfdl}
\caption{\label{fig:qsl-tbi-ener} (Color online) Energy derivative
with respect to $\lambda$ [Eq.~(\ref{eq:dfdlambda})], across the QSL--TBI transition at
$U/t=4$. There is no sign of a discontinuity at $\lambda_\text{c}/t=0.030(1)$.}
\end{figure}
\section{Conclusions and Outlook}\label{sec:conclusions}
Using exact quantum Monte Carlo simulations, we have obtained the phase
diagram of the Kane-Mele-Hubbard model (Fig.~\ref{fig:pd}). For
weak Hubbard interaction, the system is either a semimetal (SM) (at zero
spin-orbit coupling, $\lambda=0$) or a topological band insulator (TBI). The
latter is adiabatically connected to the noninteracting groundstate of the
Kane-Mele model, as evinced by the almost identical dependence of the single-particle and
spin gaps on $\lambda$. We have presented evidence for substantial
particle-hole binding in the TBI phase for small systems or large Hubbard
interaction. For intermediate Hubbard $U$, the
model supports a quantum spin liquid (QSL) phase at small $\lambda$ and a TBI phase at large
$\lambda$. At large $U$, long-range magnetic order breaks time reversal
invariance, and the system becomes an antiferromagnetic Mott insulator (AFMI).
In the presence of spin-orbit coupling, magnetic order is restricted to the
$xy$ plane.
As previously suggested,\cite{Hohenadler10,Gr.Xu.11,PhysRevLett.107.166806}
the magnetic TBI--AFMI transition can be understood
as a condensation of magnetic excitons. A scaling analysis of the
magnetization and the spin gap provides clear evidence for the 3D XY
nature of the transition. The onset of long-range order coincides
with the closing of the spin gap, whereas the single-particle gap stays
finite but shows a cusp at the critical point. In contrast to
theoretical predictions, the corresponding transition between the QSL and the
AFMI appears to be continuous.
The QSL--TBI transition manifests itself as a cusp in the single-particle
and spin gap. The numerical data are compatible with a complete closing of
the gaps, but a definite conclusion is complicated by restrictions in
lattice sizes. The independently deduced inverse spin correlation length
is consistent with this picture, thereby suggesting
that the QSL and TBI phases are not adiabatically connected. Finally, we find
no sign of a predicted first-order transition.
There remain a number of interesting open issues, including a characterization of
the QSL phase, resolving the possible closing of the spin and single-particle
gap across the QSL--TBI transition, and the universality class of the QSL--AFMI
transitions both at $\lambda=0$ and $\lambda>0$. Understanding the
universality would provide important insight about the nature of the QSL
phase, including the possible existence of fractionalization. All
these questions require significantly larger system sizes and
hence massively parallel computers and will be addressed in future work.
{\begin{acknowledgments}%
We thank G.~Fiete, A.~Ruegg, C.~Varney and C.~Xu for useful
discussions. We acknowledge support from the DFG Grants No.~FOR1162,
SFB/TRR21 and WE 3639/2-1. This research was supported in part by the
National Science Foundation under Grant No. NSF PHY0551164 and the NSF
EPSCoR Cooperative Agreement No. EPS-1003897 with additional support from
the Louisiana Board of Regents.
Z.Y.M. acknowledges the hospitality of the Institute of Physics and KITPC at
the Chinese Academy of Sciences. We are grateful to LRZ Munich, NIC
J\"ulich, the J\"ulich Supercomputing Centre and HLR Stuttgart for
generous allocation of computer time.
\end{acknowledgments}}
|
\section{Introduction}
Progress in metamaterial design, from microwaves to THz and optical
frequencies \cite{Soukoulis2002, Alu2005, Smith2005, Soukoulis2006,
Soukoulis2007}, based on the magnetic response of arrays of wires
and split ring resonators (SRR) and their modifications to shorter
electromagnetic waves, shows as essential limitations the existence
of anisotropy and large absorption losses \cite{Soukoulis2008}. More
recently, alternative structures based on Mie scattering by
dielectric spheres \cite{Soukoulis2008, Zhang2009_1, Mojahedi2005,
Mojahedi2006, Brongersma2007, Peng2007} of relative large refractive
index were extensively studied and proposed as models of lossless
composites at microwaves; these were also extended to semiconductor
cylinders and spheres, which were proven to possess similar resonant
characteristics in the infrared and visible regions \cite{Vynck2009,
Nieto2011}. Although initially these latter structures relied on
only the electric dipole or on multipole modes, later the
possibility of exciting the first Mie magnetic dipole resonance was
realized; which makes these particles equivalent from a fundamental
point of view to the microwave wire and SRR elements as far as their
electric and magnetic responses are concerned. Moreover, these
cylinders and spheres provide isotropy 2D and 3D in addition to
their resonance being subwavelength, both for the electric and
magnetic excited dipoles. The purpose of this paper is to
investigate the transmittance properties of media composed of these
dielectric particles and in particular, whether they behave as
uniform media at frequencies where their electric and magnetic
dipoles are excited. Therefore this work is a test and assessment of
whether the excitation of electric and magnetic resonances of high
index Mie spheres and cylinders constitutes an alternative with low
losses and high refractive transmittivity, to previously developed
metamaterial models.
In this connection, we emphasize that in spite of the exhaustive
studies on the possibilities of these resonant particles as
metamaterial building blocks in \cite{Mojahedi2005, Mojahedi2006,
Brongersma2007, Peng2007, Vynck2009, Nieto2011} and references
therein, no such a test has been carried out.
Although a unified view of the effective refractive index of
metamaterials made of ordered arrays of such particles, as well as
of photonic crystals (PC) producing negative refraction
\cite{Notomi2000} was established \cite{Zhang2010}, the
characterization of these composites as effective homogeneous media
to the propagating wave has generally employed the method of the
scattering (S) parameter by inversion of the complex transmittances
and reflectances \cite{Soukoulis2002}. Then, it remained the
question of whether the effective constitutive parameters
$\epsilon_{eff}$ and $\mu_{eff}$ derived from effective medium
homogenization procedures commonly employed \cite{Mojahedi2005,
Mojahedi2006, Peng2007, Nieto2011}, were the same as those obtained
by those methods that took into account the wave interaction with
the microstructure of the composite unit cell. The answer, recently
given \cite{Zhang2011}, is negative. {\it The linear dimension of
such unit cells, which is typically between $\lambda/10$ and
$\lambda/6$, is a much larger number than those of atom or molecule
arrangements in transparent dielectrics acting to the pass of
light}. In fact, the maximum lattice constant versus wavelength for
a metamaterial to behave as an uniform effective medium in negative
refraction experiments, was established in \cite{Zhang2008}.
Therefore, one may ask whether reducing absorption of the composite,
like with these Mie resonances of dielectric spheres and cylinders,
is sufficient to achieving an application as a metamaterial which
may be considered as a refractive element.
To this end, we shall see in this paper that, although contrary to
metal elements like those of earlier left - handed material (LHM)
designs, these dielectric cylinders and spheres grouped as "meta -
atoms{\lq\lq} of composites do not present absorption, their
scattering cross section is quite large. Therefore, {\it we will
demonstrate that this time the losses of light transmission come
from their large extinction cross section due to scattering, which
moreover, when random arrangements of these particles are also
addressed, yields a rather short transport mean free path, even if
the medium is homogenized, and hence the transmittivity of a
propagating beam through these media is low}.
As a 2-D equivalent to spheres, both ceramic \cite{Peng2007} and of
Silicon \cite{Vynck2009}, rod array composites were suggested to
exhibit metamaterial left-handed behaviour in the microwave and in
the visible to mid-infrared (IR) ranges, respectively, due to the
excitation of the cylinder magnetic Mie resonance. However, further
studies pointed out \cite{Zhang2009_2} the necessity of {\it more
research to short out whether the backward wave behavior inside a
periodic array of such Si cylinders is due to the band structure in
the diffraction regime \cite{Vanbesien2008} or to a pure left-handed
effect in the long wavelength range}, even though experiments of
microwaves in prisms of large permittivity ceramic rod arrays,
either ordered or random, suggested a left-handed behavior
\cite{Peng2007}.
The typical filling fraction $f$ of the studied dielectric cylinder
arrays \cite{Peng2007, Vynck2009, Nieto2011} is moderate $f\approx
0.30$, but the wavelength in vacuum to rod radius ratio: $\lambda/r$
was $61$ in e. g. the experiment of \cite{Peng2007}
($\lambda=41.64mm$), the lattice constant $a$ to $\lambda$ ratio
being 0.07, which is well below the aforementioned ratio $a/\lambda=
0.1$; whereas $\lambda/r= 9.8$ and $a/\lambda= 0.45$ in the model of
the ordered Si cylinders of \cite{Vynck2009} ($\lambda=1.55\mu m$).
On the other hand, the equivalent homogeneous media obtained from
Snell law had an index of negative refraction $n\simeq -0.6$ for the
PC slab of \cite{Vynck2009} and $n\simeq -1.08$ for the PC prism of
\cite{Peng2007}, ($\lambda_n\simeq 2.58\mu m$ for the case of
\cite{Vynck2009}, whereas $\lambda_n\simeq 38.5mm$ for
\cite{Peng2007}, $\lambda_n=\lambda/n$). None of these values is
maintained when one changes the sample geometry, as we shall prove
in this paper.
In the homogenization procedure employed in the ceramic composite of
\cite{Peng2007}, the effective parameters obtained in the band of
left-handed behaviour have negative values both for the real part of
the magnetic permeability, $\mu^R_{eff}$, as for that of the
dielectric permittivity, $\epsilon^R_{eff}$, with $\mu^R_{eff}<<
\epsilon^R_{eff}$. However, near the resonance wavelength, the
imaginary parts $\epsilon^I_{eff}$ and $\mu^I_{eff}$ of both
constitutive parameters are non - negligible compared to those real
parts; {\it this conveys high extinction by the composite material,
which is not due to absorption losses as in previous models of LHMs
composed of metallic elements, but produced by scattering from the
high index particles of these metamaterials}. The same happens in
the near IR for the composite material of Si cylinders
\cite{Vynck2009} and \cite{Nieto2011}. This will be further studied
in this work.
In this paper we carry out 2 - D numerical experiments with the
finite element method (FEM) of propagation, of mid - IR waves,
through a composite medium made of dielectric rods either in
periodic or random positions. Nonetheless, for the sake of
comprehensiveness, we shall also address the related problem of
microwave propagation at larger scales. From the above discussion,
our aim is to assess for the first time to what extent these
structures may constitute a metamaterial model that overcomes the
losses \cite{Schultz2001, Nieto2003} of previous composites. This
conveys to discuss the validity of establishing effective
constitutive parameters. The results should closely predict
laboratory experimental observations because the calculations
involved are exact. Since in this model the size of the samples are
not huge compared to the wavelength \cite{Zhang2011, Zhang2008}, in
order to assess whether the propagation depends or not on this size
and on the shape of the sample, we employ two of such bodies for
observing transmission: a rectangular block, or thick slab, and a
prism, (i. e. in the 2-D calculations the latter being a triangle),
composed of either a periodic or a disordered array of rods in air.
Like in previous studies on these composites, the incident wave is
assumed to be linearly polarized with electric vector ${\bf E}$
along the cylinder axes.
\section{Sets of ordered and random rods as metamaterials}
\label{}
\subsection{Numerical procedure}
Maxwell equations are solved by using a finite element method (FE)
(FEMLAB of COMSOL, \mbox{http://www.comsol.com}). The calculation
domain is meshed with element growth rate: 1.55, meshing curvature
factor: 0.65, approximately. The geometrical resolution parameters
consist of 25 points per boundary segment to take into account
curved geometries in order to adapt the finite elements to the
geometry and optimize the convergence of the solution. The final
mesh contains about $10^{4}$ elements. To solve Helmholtz equation,
the UMFPACK direct is employed. The boundary conditions of the
simulation space are established both to keep the calculations from
undesired window reflections and to avoid possible geometrical
discontinuities. We then ensure that no inconsistencies due to
properties discontinuities of the objets under study appear, and
possible systematic errors are avoided.
In the 2 - D configuration, light depolarization is prevented by
launching linearly polarized light with propagation vector in the XY
- plane of the cylinder cross sections. Beam profiles are either,
rectangular (plane waves): ${\bf E_0}\exp(i({\bf k_i}\cdot {\bf
r_i}-\omega t))$, their widths being that of the simulation window,
or Gaussian: ${\bf E_0}\exp(-|{\bf R}-{\bf
R_0}|^2/2\sigma^2)\exp(i({\bf k_i}\cdot {\bf r}-\omega t))$, ${\bf
r}= ({\bf R}, z)$. ${\bf R}= (x, y)$ and ${\bf R_0}= (x_0, y_0)$ are
the transversal components of ${\bf r}$ and ${\bf r_0}$,
respectively, $\sigma$ is the standard deviation or beam waist, and
${\bf k_i}$ is the incident wave wavevector with ${\bf
|k_i|}=2\pi/\lambda$. For the wavelength $\lambda$, we shall address
values either in the microwave or IR regions. The direction of
propagation ${\bf k_i}$ of such beams is thus normally incident to
the OZ - axis of the infinite cylinders. The incident wave amplitude
is normalized to $|{\bf E_0}|= 1V/m$ (SI), which corresponds to a
magnitude of the time average energy flow $|<{\bf S}>|\approx
190W/m^{2}$. The criterium to choose the beam profile has been based
on the major response of the composite characteristics to analyze:
directionality of propagation in the media resulting from an
homogenization method, and extinction processes when the inner
structure of the rod distribution is considered. In the first case,
we employ rectangular incident beams; otherwise, we use Gaussian
beams.
The results are thus expressed in terms of either the electric
vector ${\bf E}({\bf R})$ which points along the cylinder OZ - axis,
the squared root of its time - averaged energy ${\bf |E(R)|}$, the
magnetic vector ${\bf H}({\bf R})= (H_x, H_y)$ or the time average
energy flow $<{\bf S}({\bf R})>$. These latter two vectors of course
being both transversal, namely, in the XY - plane of the images to
show next.
Finally, to classify the whispering gallery modes (WGM) associated
to the Mie resonances of the cylinders we will use the subscripts
\emph{(i, j)}, \emph{i} and \emph{j} standing for their angular
\emph{i - th} and radial \emph{j - th} orders, respectively.
\subsection{Electric and magnetic dipolar response of a dielectric cylinder in the microwave and mid - IR regimes}
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=6cm]{fig1a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=6cm]{fig1b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=6cm]{fig1c}
\end{minipage}
\caption{(a) Electric field $E_z({\bf R})$ in a cylinder of BST
ceramic with a dielectric permittivity $\epsilon= 600$ and radius
$r= 0.68mm$. (b) Magnetic field ${\bf H(R)}$ (arrows) and its X -
component (colors). (c) Averaged energy flow ${\bf <S(R)>}$ (arrows)
and its norm (colors). In these figures, an s - polarized Gaussian
beam of amplitude $A= 1V/m$ and standard deviation $\sigma= 12mm$ at
$\lambda= 41.638mm$ is launched upwards (i. e. with ${\bf R_i}$
along the OY - axis), from below the cylinder, exciting its $WGM:
TM_{1,1}$.}
\end{figure}
{\noindent We first address the response of one single ceramic}
cylinder of $Ba_{0.5}Sr_{0.5}TiO_{3}$ (BST) to microwaves. Figures
1(a) - (c) show the electric, magnetic and time - averaged Poynting
vector distributions for linear polarized illumination with the
electric vector along the cylinder axis. Figure 1(a) exhibits
$E_z({\bf R})$, indicating two electric currents flowing along the
cylinder OZ - axis, one upwards and one downwards, respectively, and
centered near opposite sides of the rod periphery, which corresponds
to the {\bf E} - spatial distribution of the dipolar $WGM:
TM_{1,1}$. Figure 1(b) shows the magnetic vector lines in the XY -
plane, characterized by arrows circulating around these electric
currents, according to Ampere's law and behaves as that of a
magnetic dipole. The time - averaged energy flow, characterized by
the mean Poynting vector, is shown in Fig. 1(c) exhibiting an
interesting circulation around the equatorial extremes of the
cylinder section.
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=6cm]{fig2a}
\end{minipage}
\begin{minipage}{.49\linewidth} \centering
\includegraphics[width=6cm]{fig2b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=6cm]{fig2c}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ in a cylinder of Si with
a dielectric permittivity $\epsilon= 12$ and radius $r= 158nm$. (b)
Magnetic field ${\bf H(R)}$ (arrows) and its X - component (colors).
(c) Averaged energy flow ${\bf <S(R)>}$ (arrows) and its norm
(colors). An s - polarized Gaussian beam of amplitude $A= 1V/m$ and
standard deviation $\sigma= 2792nm$ at $\lambda= 1.55\mu m$ is
launched upwards (i. e. with ${\bf R_i}$ along the OY direction),
from below the cylinder, exciting its $WGM: TM_{1,1}$.}
\end{figure}
On the other hand, Figs. 2(a) - 2(c) show the response of one single
Si cylinder to infrared light. The characteristics quoted above for
the BST ceramic cylinder are again reproduced for this Si rod, as
expected from the analysis of their similar electric and magnetic
resonances \cite{Vynck2009}.
\subsection{Transmission characteristics of a slab with ordered rod distributions in the microwave regime}
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig3a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig3b}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig3c}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig3d}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ spatial distribution of
a microwave propagating through a slab of an ordered array of BST
rods as the one of Figs. 1(a) - (c) ($\epsilon= 600$, $r= 0.68mm$).
The lattice constant is $a= 3mm$. An s - polarized Gaussian beam of
amplitude $A= 1V/m$ and width $\sigma= 4a$ is launched on the PC
from the left at wavelength $\lambda= 41.638mm$ ($\nu= 7.2GHz$) and
at an incidence angle $\theta= 20^\circ$ with the X - axis. (b)
Electric field norm $|E_z({\bf R})|$ (colors) and ${\bf <S(R)>}$
(arrows) in a detail of the block upper left corner. (c) Same as in
(a) in a slab occupied by a homogeneous medium whose effective
parameters obtained from the EMT are $\epsilon_{eff}= -36 + i14.7$,
$\mu_{eff}= -0.9 + i0.11$. (d) Electric field norm $|E_z({\bf R})|$
(colors) and ${\bf <S(R)>}$ (arrows) in a detail of the block upper
left corner.}
\end{figure}
We now consider a thick slab made of a periodic array of the BST
rods in air whose response was studied in Figs. 1(a) - (c), (see
Figs. 3(a) and 3(b)). An s - polarized light beam is launched on
this system at angle $\theta= 20^\circ$. The map of transmitted
$E_z({\bf R})$ field is shown in Fig. 3(a). A transmission angle
$\theta\approx 0^{\circ}$ is observed within the sample. A resonant
field distribution appears within the rods, being somewhat similar
to that of Fig. 1(a), taking into account the frequency shifts due
to the presence of neighbor rods. On the other hand, Fig. 3(b) shows
both the time - averaged electric energy $|E_z({\bf R})|$ and energy
flow ${\bf <S(R)>}$ in a detail of the block of rods. The latter
showing the Bragg directions of propagation inside the crystal.
Nevertheless, in the experiment of \cite{Peng2007} (cf. Figs. 3 of
Ref. \cite{Peng2007}) on a prism of an identical array, a refraction
angle of about $20^\circ$ was obtained instead, which would
correspond to $n_{eff} \simeq -1.08$. A 75\% of transmitted energy
is lost in the rod slab of Figs. 3(a) and 3(b).
\begin{figure}[htbp]
\centering
\includegraphics[width=7cm]{fig4}
\caption{Estimation of both the complex relative permittivity
$\epsilon_{eff}$ and permeability $\mu_{eff}$ for the BST rod array
in Fig. 3(a) obtained the effective medium theory.
$\epsilon^R_{eff}$ (broken line with triangles. Blue online),
$\epsilon^I_{eff}$ (broken line with dots. Green online),
$\mu^R_{eff}$ (full line. Red online) and $\mu^I_{eff}$ (broken
line. Black online). The broken with dots vertical line (violet
online) at frequency $\nu= 7.2GHz$ indicates both $\epsilon_{eff}$
and $\mu_{eff}$ values for the medium of Figs. 3(c) and 3(d):
$\epsilon_{eff}= -36$, $\epsilon^I_{eff}= 14.7$, $\mu^R_{eff}= -0.9$
and $\mu^I_{eff}= 0.11$. The full vertical (blue online) line
belongs to the jump of $\epsilon^R_{eff}$.}
\end{figure}
However, homogenization from an effective medium theory employed
like in \cite{Peng2007, Nieto2011, Lewin1948, Holloway2003} leads us
to estimate for this composite the effective homogeneous medium
constitutive parameters $\epsilon^R_{eff}= -36$, $\epsilon^I_{eff}=
14.7$; $\mu^R_{eff}= -0.9$, $\mu^I_{eff}= 0.11$ (see Fig. 4). Such a
uniform medium produces an angle of refraction which is very small,
similar to that of Fig. 3(a) but does not reproduce the inner
structure of the wavefield. This is shown in Figs. 3(c) and 3(d)
which display $E_z({\bf R})$ and a detail of $|E_z({\bf R})|$ and
${\bf <S(R)>}$ for such an effective medium.
Now a 95\% of losses appears in the transmitted energy emerging from
the homogeneous medium of Figs. 3(c) and 3(d). These results
manifest the very different observed refraction depending on the
sample on use: whether it is the composite or we deal with its EMT
homogenization. In fact, although not shown here, we remark that we
have reproduced the numerical results of \cite{Peng2007} for such an
array if the sample is a prism which, as said above, yields a
refraction completely different to that of Figs. 3(a) and 3(b).
Hence, we infer the questionable matching of an EMT with the
observations in such a regular array of rods. This is not surprising
in the light of the results of \cite{Zhang2011} and
\cite{Zhang2008}, since the sample in these cases is not much larger
than the inner structure, at difference with light transmission
experiments in usual light refractive elements of nanoscopic
molecular inner structure sizes.
\subsection{Microwave transmission in a slab with a random distribution of cylinders}
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig5a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig5b}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ spatial distribution in
a random configuration of cylinders like those of Figs. 1(a) - (c)
keeping the same filling fraction and the same conditions of
illumination as in the ordered array of Figs. 3. (b) Electric field
norm $|E_z({\bf R})|$ (colors) and ${\bf <S(R)>}$ (arrows) in a
detail of the block central region which includes its upper
transparent side.}
\end{figure}
{\noindent Different realizations, obtained by randomizing the}
array of Figs. 3(a) and 3(b) are now addressed. We next study a
disordered distribution of BST rods as those employed in Section
~\ref{ORS_MW}. Now, there is 85\% of transmission losses through the
slab of these disordered rods due to extinction produced by
scattering. Figs. 5(a) and 5(b) show the field $E_z({\bf R})$ and a
detail of both $|E({\bf R})|$ and ${\bf <S(R)>}$, respectively. A
large amount of scattered light is lost both above and below the
slab upper and lower low reflection boundaries. Again some few rods
exhibit the excitation of Mie resonances on illumination as in Fig.
1(a). Figures 5(a) and 5(b) yield no clue of a forward transmission
direction into the air at the exit of the sample. This result is
obtained with both a beam and a plane wave. Although not shown here,
we should state that we observed that averaging the field $E_z({\bf
R})$ over many realizations of the random distribution of rods does
not yield a forwardly transmitted beam, characterized by
$<E_{z}({\bf R})>$ \cite{Ishimaru, Nieto1997, Nieto2000, Nieto2001},
distinguishable from the scattered light.
\subsection{Ordered and random distributions of Silicon rods in the mid - infrared regime}
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig6a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig6b}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig6c}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig6d}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ propagating in a thick
slab occupied by an ordered array of Si rods ($\epsilon= 12$ and $r=
158nm$; the lattice constant is $a= 698nm$). An s - polarized
Gaussian beam of amplitude $A= 1V/m$, $\sigma= 4a$ and wavelength
$\lambda= 1.55\mu m$ is launched on the slab from the left at an
incidence angle $\theta= 20^\circ$. (b) Electric field norm
$|E_z({\bf R})|$ (colors) and ${\bf <S(R)>}$ (arrows) in a detail of
the block upper left corner. (c) Electric field $E_z({\bf R})$ in a
uniform medium optically equivalent to that of (a) and (b), whose
electric permittivity is $\epsilon_{eff}= -0.36 + i0.25/8$. The real
part of such a value is estimated by Snell's law, ($\theta_{i}=
21.54^\circ$, $\theta_{t}= 38.07^\circ$). The imaginary part has
been estimated from the transmittivity of Figs. 6(a) and 6(b). (d)
Electric field norm $|E_z({\bf R})|$ (colors) and ${\bf <S(R)>}$
(arrows) in a detail of the block upper left corner.}
\end{figure}
We next address a thick slab of ordered Si cylinders like that
studied in Figs. 2. This arrangement is an extension to mid - IR
(\cite{ Vynck2009}) of the previous calculations at microwaves
discussed in Sections ~\ref{ORS_MW} and ~\ref{RRS_MW}. As shown in
Figs. 6(a) and 6(b) there appears negative refraction (i. e. a
backward wave) in the block and no appreciable transmission losses.
The positive and negative peak values of the wavefronts transmitted
inside the array coincide with those cylinders that appear
resonantly illuminated thus exhibiting the Mie $T_{1, 1}$ resonance
like in Fig. 2(a). Mainly, there are ${\bf <S(R)>}$ arrows pointing
in the direction normal to these wavefronts. Light is transmitted
into the air side on the right through this slab with the same angle
as that of incidence. This confirms the analysis of \cite{Vynck2009}
for this crystal.
On the other hand, Figs. 6(c) and 6(d) display the transmission of
the same incident wave through a thick slab occupied by a uniform
medium with a refractive index: $n= -0.36 + i0.25/8$ which resembles
the propagation through the ordered array shown in Figs. 6(a) and
6(b). The real part $n^R= -0.36$ has been estimated from Snell's law
applied to Fig. 6(a), whereas the imaginary part $n^I= 0.25/8$ was
fitted to a transmittance of this homogeneous medium being
approximated to that of Figs. 6(a) and 6(b). Notice that the
transmittivity of the homogeneous slab of Figs. 6(c) and 6(d) is
smaller than that of the ordered array of Figs. 6(a) and 6(b).
However $n^I$ cannot decrease much beyond 0.25/8, since otherwise
well known instabilities due to divergences at the right side of the
slab \cite{Heyman2001, Nieto2002, Nieto2004} appear. However, it
should be remarked that this value of $n$ does not coincide with
that of an EMT, which according to the values of $a/\lambda$ and
$r/\lambda$ for this array, as discussed in Section ~\ref{Intro},
are too large for a homogenization procedure to work with such a
structure. This is further discussed next by employing other arrays
of these cylinders with the same filling fraction and lattice
parameter; namely, a thick slab of disordered Si rods and a prism of
ordered, or disordered, cylinders.
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig7a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig7b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=7cm]{fig7c}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ when the configuration
of Si rods shown in Figs. 6(a) and 6(b) is disordered, keeping the
same filling fraction $f= 0.30$ and the same conditions of
illumination. (b) Electric field norm $|E_z({\bf R})|$ (colors) and
${\bf <S(R)>}$ (arrows) in an inset of the block left region which
includes its left interface. (c) Map of the averaged energy flow
norm ${\bf |<S(R)>|}$.}
\end{figure}
When these cylinders are disordered keeping $f\approx0.30$, as seen
in Figs. 7(a) - (c), there is no such negative refraction inside the
block as that shown in Figs. 6. Now there is a huge extinction of
energy due to high scattering by the particles. By averaging over
several realizations, there is no observed refractive transmission
of a forward component $<{\bf E_{z}}>$ into the air side. The block
of disordered rods now scatters, see Figs. 7(a) and 7(b), more than
90\% of the transmitted intensity both above and below its upper and
lower low reflection boundaries. Hence, only some few spheres are
illuminated showing their $TM_{1, 1}$ Mie resonances. Figure 7(c)
shows that ${\bf |<S(R)>|}$ is transmitted into the air region at
the right of the slab with less than a 1/10 of its incident wave
value.
\subsection{A prism of ordered or disordered Si distributions in the infrared}
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig8a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig8b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=7cm]{fig8c}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ in a prism (top angle
$\alpha= 18.4^\circ$) occupied by a uniform medium whose electric
permittivity is $\epsilon_{eff}= -0.36 + i0.25/8$ illuminated from
the left. This prism is optically as dense as the homogeneous thick
slab of Figs. 6(c) and 6(d). (b) Electric field norm $|E_z({\bf
R})|$ (colors) and ${\bf <S(R)>}$ (arrows) in a detail of the prism
upper region. (c) Map of the averaged energy flow norm ${\bf
|<S(R)>|}$. An s - polarized Gaussian beam of amplitude $A= 1V/m$,
$\sigma= 4\times 698nm$ and wavelength $\lambda= 1.55\mu m$ is
launched from the left on the prism at $\theta_i= 0^\circ$ with its
left side.}
\end{figure}
{\noindent Let us consider now a prism of a uniform medium of}
refractive index $n= -0.36 + i0.25/8$. Notice that this value of $n$
is the one of Figs. 6(c) and 6(d), derived from refraction at the
ordered array of Figs. 6(a) and 6(b). As shown in Figs. 8,
refraction into the air takes place with $\theta_{i}= 18.4^{\circ}$
and $\theta_{t}= 38.07^{\circ}$, the latter being the negative angle
of refraction at the prism larger side. This is shown in Figs. 8(a)
- (c) and serves us as a reference to study transmission through
both an ordered and disordered array of Si rods contained in a
sample with this prism geometry, which also was the one employed in
transmission observations at microwaves in \cite{Peng2007}.
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig9a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig9b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=7cm]{fig9c}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ in a prism with the same
shape as that of Figs. 8(a) - (c), at the same illumination
conditions, but now occupied by an ordered Si rod array as that of
Figs. 6(a) and 6(b). (b) Electric field norm $|E_z({\bf R})|$
(colors) and ${\bf <S(R)>}$ (arrows) in a detail of the prism upper
region. (c) Averaged energy flow norm ${\bf |<S(R)>|}$. The
illumination is the same as in Figs. 8(a) - 8(c).}
\end{figure}
Figures 9(a) - (c) show the wave propagation on illumination of the
prism filled with an ordered array of Si cylinders identical to that
of the thick slab of Figs. 6(a) and 6(b). As shown, there is now
absence of negative refraction at the larger side interface of this
prism. This contrasts with the observation in the sample with the
same shape filled with a uniform medium, as displayed in Figs. 8(a)
- (c) and also with the case of a thick slab filled with the same
array, as seen in Figs. 6(a) and 6(b). Hence, at difference with the
crystal of Si rods in the block, the same crystal in the prism does
not reproduce negative refraction, but rather a set of diffracted
orders into the air which are associated to the prism angle
$\alpha$, according to the conservation of the transversal
wavevectors at the larger interface. Also, the wave propagation
inside the prism, which should be like that appearing in the thick
slab of Fig. 6(a), but now at the same direction as the incident
wave, (since now this latter wave incides on the prism at $\theta_i=
0^\circ$), is observed in Figs. 9(a) and 9(b) to be quite different,
with a complicated structure due to the interference of the
different Bragg waves. This result, once again, points out the
dependence of wave propagation not only on the inner structure of
the composite, but also on its shape. In addition, these
observations demonstrate that the negative refraction found in the
slab of ordered Si rods, (cf. Fig. 6(a)) is not due to an effective
homogeneous medium effect, but it rather comes from a consequence of
the diffraction in the array of Si rods due to its lattice symmetry.
This answers the question posed in \cite{Zhang2009_2}.
\begin{figure}[htbp]
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig10a}
\end{minipage}
\begin{minipage}{.49\linewidth}
\centering
\includegraphics[width=7cm]{fig10b}
\end{minipage}
\begin{minipage}{.98\linewidth}
\centering
\includegraphics[width=7cm]{fig10c}
\end{minipage}
\caption{ (a) Electric field $E_z({\bf R})$ in the configuration
shown in Figs. 9(a) - (c), (now with a top angle $\alpha=
22^\circ$), the Si rod array being now randomized but keeping the
same filling fraction $f= 0.30$ as in the ordered array. The
conditions of illumination are the same as in Figs. 8(a) - 8(c). (b)
Electric field norm $|E_z({\bf R})|$ (colors) and ${\bf <S(R)>}$
(arrows) in a detail of the prism upper area. (c) Averaged energy
flow norm ${\bf |<S(R)>|}$.}
\end{figure}
This latter remark is further confirmed by disordering this Si
cylinder array within the prism, (see Figs. 10(a) - (c)). Again, no
refracted beam into the air is observed, and a large portion of
energy is lost by scattering from the composite elements. Many
outgoing beams mix with each other, so that no forwardly transmitted
beam inside the prism, characterized by $<E_{z}({\bf R})>$ can be
distinguished by averaging over several realizations of the random
array. The apparent refracted beam at $\theta_t= 0^\circ$ shown in
the air side exiting the prism, changes as one varies the random
realization of rods.
\section{Conclusions}
In this paper we have assessed the transmittance of composites of
dielectric particles whose Mie resonance electric and magnetic modes
were proposed by previous extensive studies to yield negative
refraction. We have shown that such structures cannot be homogenized
neither reproduce the propagation observed in the frequency regions
of study. In this way, we have proved that the negative refraction
previously found in ordered arrays is thus a diffraction effect
which disappears as soon as the particle distribution is randomized,
and does not reproduce the transmission of their corresponding EMT
uniform media, this was an open question so far.
In addition, we conclude, first, that the effective parameters
obtained from a homogenization theory, do not reproduce the
propagation observed through an ordered array of these elements.
This is further seen when these arrays are randomized. If an EMT
worked, it should not depend on whether the \lq\lq meta - atom"
distributions were ordered or disordered. Then strong scattering by
disordered rods extinguishes most of the incident energy, and there
is not negatively refracted forward beam observed.
Second, the behavior of the transmitted beam in ordered arrays also
depends on the shape of the sample and hence does not match with an
EMT which does not include this shape. These conclusions are
consistent with the well known {\it difficulty of working out EMTs
within the frequency range of resonance of the composite elements}
\cite{Zhang2011}. Similar consequences should hold for 3D composites
of resonant dielectric spheres.
\section*{Acknowledgements}
We thank J. J. S\'aenz and L. Froufe for stimulating and useful
discussions on this subject. Research supported by the Spanish
MiCINN through FIS2009-13430-C02-C01 and Consolider NanoLight
(CSD2007-00046) research contracts. The latter grant supports the
work of FJVV.
|
\section{Introduction}
The classical McKay correspondence appears in various contexts. It relates finite subgroups $\Gamma$ of $SU(2)$ with the algebraic geometry of the quotient Kleinian singularities $\mathbb{C}^2/\Gamma$ \cite{reid:2002} but also with the classification of $SU(2)$ modular invariants
and quantum subgroups of $SU(2)$ \cite{cappelli/itzykson/zuber:1987ii, zuber:2002, ocneanu:2000ii, ocneanu:2002, xu:1998, bockenhauer/evans:1999i, bockenhauer/evans:1999ii, bockenhauer/evans/kawahigashi:1999, bockenhauer/evans/kawahigashi:2000}.
Minimal resolutions of Kleinian singularities can be described via the moduli space of representations of the preprojective algebra associated to the action of $\Gamma$ \cite{crawley-boevey/holland:1998}.
Preprojective algebras associated to graphs were introduced in \cite{gelfand/ponomarev:1979}, and it was shown that they are finite dimensional if and only if the graphs are of $ADE$ type, that is, one of the simply laced Dynkin diagrams.
The preprojective algebra $A$ for an $ADE$ Dynkin diagram is a Frobenius algebra, that is, there is a linear function $f:A \rightarrow \mathbb{C}$ such that $(x,y):=f(xy)$ is a non-degenerate bilinear form (this is equivalent to the statement that $A$ is isomorphic to its dual $\widehat{A} = \mathrm{Hom}(A,\mathbb{C})$ as left (or right) $A$-modules). There is an automorphism $\beta$ of $A$, called the Nakayama automorphism of $A$ (associated to $f$), such that $(x,y) = (y,\beta(x))$, which yields an $A$-$A$ bimodule isomorphism $\widehat{A} \rightarrow {}_1 A_{\beta}$ \cite{yamagata:1996}.
The Nakayama automorphism for each $ADE$ graph was determined in \cite{erdmann/snashall:1998i, erdmann/snashall:1998ii} (see also \cite{brenner/butler/king:2002}).
The preprojective algebra $A$ has a finite resolution as an $A$-$A$ bimodule, which was used by Erdmann and Snashall to determine the Hochschild cohomology $HH^{\bullet}(A)$ of $A$ for the $ADE$ graphs $A_n$, along with its ring structure \cite{erdmann/snashall:1998i}, and $HH^2(A)$ for the graphs $D_n$ \cite{erdmann/snashall:1998ii}
This finite resolution yields a projective resolution of $A$ as an $A$-$A$ bimodule, which was used by Etingof and Eu to determine the Hochschild homology and cohomology, and cyclic homology, of $A$ for all $ADE$ Dynkin diagrams \cite{etingof/eu:2007}, along with the ring structure of the Hochschild cohomology \cite{eu:2007i}. The Hochschild homology and cohomology, cyclic homology, and ring structure of the Hochschild cohomology, for the preprojective algebra for the tadpole graphs $T_n$ were obtained in \cite{eu:2007ii}.
The Hochschild homology and cohomology for the case of the affine Dynkin diagrams, which are the McKay graphs for the finite subgroups of $SU(2)$, were determined in \cite{crawley-boevey/etingof/ginzburg:2007}.
More generally, one tries to understand singularities via a noncommutative algebra $A$, often called a noncommutative resolution, whose centre corresponds to the coordinate ring of the singularity \cite{vandenBergh:2004}. The algebra should be finitely generated over its centre, and the desired favourable resolution is the moduli space of representations of $A$, whose category of finitely generated modules is derived equivalent to the category of coherent sheaves of the resolution.
In the case of a quotient singularity $\mathbb{C}^3/\Gamma$ for a finite subgroup $\Gamma$ of $SU(3)$, the corresponding noncommutative algebra $A$ is a Calabi-Yau algebra of dimension 3.
Calabi-Yau algebras arise naturally in the study of Calabi-Yau manifolds, providing a noncommutative version of conventional Calabi-Yau geometry.
An algebra $A$ is Calabi-Yau of dimension $n$ if the bounded derived category of the abelian category of finite dimensional $A$-modules is a Calabi-Yau category of dimension $n$. In this case the global dimension of $A$ is $n$ \cite{bocklandt:2008}.
The derived category of coherent sheaves over an $n$-dimensional Calabi-Yau manifold is a Calabi-Yau category of dimension $n$ and they appear naturally in the study of boundary conditions of the $B$-model in superstring theory over the manifold. For more on Calabi-Yau algebras, see e.g. \cite{bocklandt:2008, ginzburg:2006}.
In \cite[Remark 4.5.7]{ginzburg:2006} Ginzburg introduced, in his terminology, $q$-deformed Calabi-Yau algebras. In the case where $q$ is not a root of unity, these algebras are Calabi-Yau algebras of dimension 3.
We study these algebras in the case where $q$ is a root of unity, which are the $SU(3)$ generalizations of preprojective algebras for the Coxeter-Dynkin diagrams $ADE$. We call these algebras \emph{almost Calabi-Yau algebras}. In a recent work \cite{evans/pugh:2010ii}, we determined the Nakayama automorphism for each $\mathcal{ADE}$ graph, and constructed a finite resolution of $A$ as an $A$-$A$ bimodule, see (\ref{exact_seq-almostCY}).
Our interest in these almost Calabi-Yau algebras came from subfactor theory, and in particular, braided subfactors of von Neumann algebras, which provide a framework for studying two dimensional conformal field theories and their modular invariant partition functions.
In the case of Wess-Zumino-Witten models associated to $SU(n)$ at level $k$, the Verlinde algebra is a non-degenerately braided system of endomorphisms ${}_N \mathcal{X}_N$, labelled by the positive energy representations of the loop group of $SU(n)_k$ on a type $\mathrm{III}_1$ factor $N$, with fusion rules $\lambda \mu = \bigoplus_{\nu} N_{\lambda \nu}^{\mu} \nu$ which exactly match those of the positive energy representations \cite{wassermann:1998}. The fusion matrices $N_{\lambda} = [N_{\rho \lambda}^{\sigma}]_{\rho,\sigma}$ are a family of commuting normal matrices which give a representation themselves of the fusion rules of the positive energy representations of the loop group of $SU(n)_k$, $N_{\lambda} N_{\mu} = \sum_{\nu} N_{\lambda \nu}^{\mu} N_{\nu}$.
This family $\{ N_{\lambda} \}$ of fusion matrices can be simultaneously diagonalised:
$$N_{\lambda} = \sum_{\sigma} \frac{S_{\sigma, \lambda}}{S_{\sigma,0}} S_{\sigma} S_{\sigma}^{\ast},$$
where $0$ is the trivial representation, and the eigenvalues $S_{\sigma, \lambda}/S_{\sigma,0}$ and eigenvectors $S_{\sigma} = [S_{\sigma, \mu}]_{\mu}$ are described by the statistics $S$ matrix.
The key structure in the conformal field theory is the modular invariant partition function $Z$. In the subfactor setting this is realised by
a braided subfactor $N \subset M$ where trivial (or permutation) invariants in the ambient factor $M$ when restricted to $N$ yield $Z$. This would mean that the dual canonical endomorphism is in $\Sigma({}_N \mathcal{X}_N)$, i.e. decomposes as a finite linear combination of endomorphisms in ${}_N \mathcal{X}_N$.
Indeed if this is the case for the inclusion $N \subset M$, then the process of $\alpha$-induction allows us to analyse the modular invariant,
providing two extensions of $\lambda$ on $N$ to endomorphisms $\alpha^{\pm}_{\lambda}$ of $M$, such that the matrix $Z_{\lambda,\mu} = \langle \alpha_{\lambda}^+, \alpha_{\mu}^- \rangle$ is a modular invariant \cite{bockenhauer/evans/kawahigashi:1999, bockenhauer/evans:2000, evans:2003}.
The action of the system ${}_N \mathcal{X}_N$ on the $N$-$M$ sectors ${}_N \mathcal{X}_M$ produces a \emph{nimrep} (non-negative matrix integer representation of the fusion rules) $G_{\lambda} G_{\mu} = \sum_{\nu} N_{\lambda \nu}^{\mu} G_{\nu}$,
whose spectrum reproduces exactly the diagonal part of the modular invariant, i.e.
$$G_{\lambda} = \sum_{\sigma} \frac{S_{\sigma,\lambda}}{S_{\sigma,0}} \psi_{\sigma} \psi_{\sigma}^{\ast},$$
with the spectrum of $G_{\lambda}$ given by
$G_{\lambda} = \{ S_{\mu, \lambda}/S_{\mu,0} \textrm{ with multiplicity } Z_{\mu,\mu} \}$
\cite[Theorem 4.16]{bockenhauer/evans/kawahigashi:2000}.
The systems ${}_N \mathcal{X}_N$, ${}_N \mathcal{X}_M$, ${}_M \mathcal{X}_M$ are (the irreducible objects of) tensor categories of endomorphisms with the Hom-spaces as their morphisms. Thus ${}_N \mathcal{X}_N$ gives a braided modular tensor category, and ${}_N \mathcal{X}_M$ a module category.
In our work we have focused on braided subfactors associated to $SU(3)$ modular invariants, which are labeled by a family of graphs which we call the $SU(3)$ $\mathcal{ADE}$ graphs. The complete list of the $SU(3)$ $\mathcal{ADE}$ graphs are illustrated in \cite[Figures 5-9]{evans/pugh:2011}.
For positive integer $k < \infty$ we have a braided modular tensor category ${}_N \mathcal{X}_N = \{ \lambda_{(p,l)} | \; 0 \leq p,l,p+l \leq k \}$, a non-degenerately braided system of endomorphisms on a type $\mathrm{III}_1$ factor $N$, which is generated by $\rho = \lambda_{(1,0)}$ and its conjugate $\overline{\rho} = \lambda_{(0,1)}$, where the irreducible endomorphisms $\lambda_{(p,l)}$ satisfy the fusion rules of $SU(3)_k$:
\begin{equation} \label{eqn:fusion_ruleSU(3)}
\lambda_{(p,l)} \otimes \rho \cong \lambda_{(p,l-1)} \oplus \lambda_{(p-1,l+1)} \oplus \lambda_{(p+1,l)},
\quad
\lambda_{(p,l)} \otimes \overline{\rho} \cong \lambda_{(p-1,l)} \oplus \lambda_{(p+1,l-1)} \oplus \lambda_{(p,l+1)},
\end{equation}
where $\lambda_{(p',l')}$ is understood to be zero if $p'<0$, $l'<0$ or $p'+l' \geq k+1$.
Then a pair $(\mathcal{G},W)$, of a cell system $W$ (see Section \ref{sect:almostCYalg}) on an $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$ with Coxeter number $k+3$ yields a braided subfactor $N \subset M$ and a module category ${}_N \mathcal{X}_M$, where the associated modular invariant, labeled by $\mathcal{G}$, is at level $k$.
For such a braided subfactor, the almost Calabi-Yau algebra can be constructed via a monoidal functor $F$, which is essentially the module category ${}_N \mathcal{X}_M$, from the $A_2$-Temperley-Lieb category to the category $\mathrm{Fun}({}_N \mathcal{X}_M,{}_N \mathcal{X}_M)$ of additive functors from ${}_N \mathcal{X}_M$ to itself.
The $A_2$-Temperley-Lieb category constructed in \cite{evans/pugh:2010ii} used ideas from planar algebras, and in particular, the $A_2$-planar algebras of \cite{evans/pugh:2009iii} (see also an earlier construction of the $A_2$-Temperley-Lieb category in \cite{cooper:2007}, and of the Temperley-Lieb category in \cite{turaev:1994, yamagami:2003}).
\begin{figure}[bt]
\begin{center}
\includegraphics[width=40mm]{fig-A2-webs}\\
\caption{$A_2$ webs}\label{fig:A2-webs}
\end{center}
\end{figure}
For $m_i,n_i \geq 0$, an $A_2$-$(m_2,n_2),(m_1,n_1)$-tangle $T$ is a tangle on an rectangle with $m_2+n_2$, $m_1+n_1$ vertices along the top, bottom edges respectively, generated by $A_2$ webs (see Figure \ref{fig:A2-webs}) such that every free end of $T$ is attached to a vertex along the top or bottom of the rectangle in a way that respects the orientation of the strings, every vertex has a string attached to it, and the tangle contains no elliptic faces. We call a vertex a source vertex if the string attached to it has orientation away from the vertex. Similarly, a sink vertex will be a vertex where the string attached has orientation towards the vertex. Along the top, bottom edge the first $m_i$ vertices are source, sink vertices respectively, and the last $n_i$ are sink, source vertices respectively.
Let $V^{A_2}_{(m_2,n_2),(m_1,n_1)}$ be the quotient of the free vector space over $\mathbb{C}$ with basis the $A_2$-$(m_2,n_2),(m_1,n_1)$-tangles, by the Kuperberg ideal generated by the Kuperberg relations K1-K3 \cite{kuperberg:1996}.
Then at level $k$, the $A_2$-Temperley-Lieb category is defined to be a quotient of the category $A_2\textrm{-}TL = \mathrm{Mat}(C^{A_2})$ by the negligible morphisms, where $C^{A_2}$ is the tensor category whose objects are projections in $V^{A_2}_{(m,n),(m,n)}$ and whose morphisms are $\mathrm{Hom}(p_1,p_2) = p_2 V^{A_2}_{(m_2,n_2),(m_1,n_1)} p_1$, for projections $p_i \in V^{A_2}_{(m_i,n_i),(m_i,n_i)}$, $i=1,2$. We write $A_2\textrm{-}TL_{(m,n)} = V^{A_2}_{(m,n),(m,n)}$, and $\rho$, $\overline{\rho}$ for the identity projections in $A_2\textrm{-}TL_{(1,0)}$, $A_2\textrm{-}TL_{(0,1)}$ respectively consisting of a single string with orientation downwards, upwards respectively. Then the identity diagram in $A_2\textrm{-}TL_{(m,n)}$, given by $m+n$ vertical strings where the first $m$ strings have downwards orientation and the next $n$ have upwards orientation, is expressed as $\rho^m \overline{\rho}^n$. It is a linear combination of simple projections $f_{(i,j)}$ for $i,j \geq 0$, $0 \leq i+j < m+n$ such that $i-j \cong m-n \textrm{ mod } 3$, and a simple projection $f_{(m,n)}$, where $f_{(1,0)} = \rho$, $f_{(0,1)} = \overline{\rho}$ and $f_{(0,0)}$ is the empty diagram. The morphisms $\mathfrak{f}_{(p,l)} = \mathrm{id}_{f_{(p,l)}}$ are generalized Jones-Wenzl projections.
The $f_{(p,l)}$ satisfy the fusion rules for $SU(3)$ \cite{evans/pugh:2010ii}:
\begin{equation} \label{eqn:fusion_rule-f(k,l)}
f_{(p,l)} \otimes \rho \cong f_{(p,l-1)} \oplus f_{(p-1,l+1)} \oplus f_{(p+1,l)},
\quad
f_{(p,l)} \otimes \overline{\rho} \cong f_{(p-1,l)} \oplus f_{(p+1,l-1)} \oplus f_{(p,l+1)}.
\end{equation}
At level $k$, the negligible morphisms are the ideal $\langle \mathfrak{f}_{(p,l)} | p+l=k+1 \rangle$ generated by $\mathfrak{f}_{(p,l)}$ such that $p+l=k+1$.
The $A_2$-Temperley-Lieb category is the quotient $A_2\textrm{-}TL^{(k)} := A_2\textrm{-}TL/ \langle \mathfrak{f}_{(p,l)} | p+l=k+1 \rangle$, which is semisimple with simple objects $f_{(p,l)}$, $p,l \geq 0$ such that $p+l \leq k$ which satisfy the fusion rules (\ref{eqn:fusion_ruleSU(3)}) of $SU(3)_k$, that is we have (\ref{eqn:fusion_rule-f(k,l)}) where $f_{(p',l')}$ is understood to be zero if $p'<0$, $l'<0$ or $p'+l' \geq k+1$.
The $A_2$-Temperley-Lieb category $A_2\textrm{-}TL^{(k)}$ may be identified with the braided modular tensor category ${}_N \mathcal{X}_N$, where the object $f_{(p,l)} \in A_2\textrm{-}TL^{(k)}$ is identified with $\lambda_{(p,l)} \in {}_N \mathcal{X}_N$.
Then the monoidal functor $F$ is given on the simple objects $f_{(p,l)}$ of $A_2\textrm{-}TL^{(k)}$ by
\begin{equation} \label{eqn:functorF}
F(f_{(p,l)}) = \bigoplus_{i,j \in \mathcal{G}_0} G_{\lambda_{(p,l)}}(i,j) \, \mathbb{C}_{i,j},
\end{equation}
where
$\mathbb{C}_{i,j}$ are 1-dimensional $R$-$R$ bimodules, where $R = (\mathbb{C}\mathcal{G})_0$. The category of $R$-$R$ bimodules has a natural monoidal structure given by $\otimes_R$.
The functor $F$ is defined on the morphisms of $A_2\textrm{-}TL^{(k)}$ using the cell system $W$ and the Perron-Frobenius eigenvector of $\mathcal{G}$, see \cite[Section 2.9]{evans/pugh:2010ii}.
If $\mathcal{G}^{\mathrm{op}}$ denotes the opposite graph of $\mathcal{G}$ obtained by reversing the orientation of every edge of $\mathcal{G}$, we have that $F(\rho^m \overline{\rho}^n)$ is the $R$-$R$ bimodule with basis given by all paths of length $m+n$ on
$\mathcal{G}$, $\mathcal{G}^{\mathrm{op}}$, where the first $m$ edges are on $\mathcal{G}$ and the last $n$ edges are on $\mathcal{G}^{\mathrm{op}}$.
In particular $F(\rho^m) = (\mathbb{C}\mathcal{G})_m$, so that we have the graded algebra $\bigoplus_m F(\rho^m) = (\mathbb{C}\mathcal{G})$, the path algebra of $\mathcal{G}$.
The endomorphisms $\rho^m$ are not irreducible however, but decompose into direct sums of the generalized Jones-Wenzl projections $f_{(p,0)}$.
The natural algebra to consider is thus the graded algebra $\Sigma = \bigoplus_j F(f_{(j,0)})$, where the $p^{\mathrm{th}}$ graded part is $\Sigma_p = F(f_{(p,0)})$. The multiplication $\mu$ is defined by $\mu_{p,l} = F(\mathfrak{f}_{(p+l,0)}): \Sigma_p \otimes_R \Sigma_l \rightarrow \Sigma_{p+l}$, where $\mathfrak{f}_{(p,l)} = \mathrm{id}_{f_{(p,l)}}$.
The graded algebra $\Sigma$ is isomorphic to the almost Calabi-Yau algebra $A = A(\mathcal{G},W)$ \cite{cooper:2007, evans/pugh:2010ii}.
In Section \ref{sect:almostCYalg} we introduce the almost Calabi-Yau algebra $A = A(\mathcal{G},W)$ for a pair $(\mathcal{G},W)$ of a cell system $W$ on an $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$. Then in Section \ref{sect:resolution_almostCYalg} we determine a periodic projective resolution of $A$ as an $A$-$A$ bimodule, starting from the finite resolution of $A$ determined in \cite[Theorem 5.1]{evans/pugh:2010ii}, which will be used to determine the Hochschild (co)homology and cyclic homology of $A$ in Sections \ref{sect:Hoch_hom}-\ref{sect:Hoch_cohom}.
In Section \ref{sect:Hoch_hom-complex} we use the projective resolution determined in Section \ref{sect:resolution_almostCYalg} to construct a Hochschild homology complex for $A$, and introduce the cyclic homology of $A$ in Section \ref{sect:cyclic_hom}. We then determine the Hochschild and cyclic homology of $A$ in Sections \ref{sect:HH_0}-\ref{Sect:HH(A)-non-trivial_beta} for the graphs $\mathcal{A}^{(n)}$, $n = 4,5,6,7$, $\mathcal{D}^{(3k+3)}$, $k \geq 1$, $\mathcal{A}^{(n)\ast}$, $n \geq 5$, $\mathcal{D}^{(3k)\ast}$, $k \geq 2$, $\mathcal{E}^{(8)}$ and $\mathcal{E}^{(8)\ast}$.
Finally in Section \ref{sect:Hoch_cohom} we construct a Hochschild cohomology complex for $A$ and use this to determine the Hochschild cohomology of $A$ in the cases listed above.
The Hochschild (co)homology and cyclic homology of $A$ can be regarded as invariants for the braided subfactors associated to the $SU(3)$ modular invariants.
Beginning with a pair $(\mathcal{G},W)$ given by a cell system $W$ on an $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$, we construct a braided subfactor $N \subset M$ which yields a nimrep which recovers the graph $\mathcal{G}$ as described above. Then we can construct the algebra $A(\mathcal{G},W)$ whose Hochschild (co)homology and cyclic homology only depends on the original pair $(\mathcal{G},W)$, or equivalently, on the braided subfactor $N \subset M$.
\section{Almost Calabi-Yau algebras} \label{sect:almostCYalg}
Let $\mathcal{G}$ be a finite directed graph, and denote by $\mathcal{G}_n$ the set of all paths on $\mathcal{G}$ of length $n$. The vertices of $\mathcal{G}$ are the paths of length 0.
If $a \in \mathcal{G}_1$ is an edge on $\mathcal{G}$, we denote by $\widetilde{a} \in \mathcal{G}^{\mathrm{op}}_1$ the corresponding edge with opposite orientation on $\mathcal{G}^{\mathrm{op}}$.
The path algebra $\mathbb{C}\mathcal{G} = \bigoplus_{k=0}^{\infty} (\mathbb{C}\mathcal{G})_k$ is the graded complex vector space with basis of the $k^{\mathrm{th}}$-graded part $(\mathbb{C}\mathcal{G})_k$ given by $\mathcal{G}_k$, where paths may begin at any vertex of $\mathcal{G}$. Multiplication of two paths $a \in (\mathbb{C}\mathcal{G})_k$ and $b \in (\mathbb{C}\mathcal{G})_l$ is given by concatenation of paths $a \cdot b \in (\mathbb{C}\mathcal{G})_{k+l}$ (or simply $ab$), with $ab$ defined to be zero if $r(a) \neq s(b)$, where $s(a)$, $r(a)$ denotes the source, range vertex respectively of the path $a$.
The commutator quotient $\mathbb{C}\mathcal{G} / [\mathbb{C}\mathcal{G}, \mathbb{C}\mathcal{G}]$ may be identified, up to cyclic permutation of the arrows, with the vector space spanned by cyclic paths in $\mathcal{G}$.
Let $\partial_a : \mathbb{C}\mathcal{G} / [\mathbb{C}\mathcal{G}, \mathbb{C}\mathcal{G}] \rightarrow \mathbb{C}\mathcal{G}$ be the derivation given by
$\partial_{a} (a_1 \cdots a_n) = \sum_{j} a_{j+1} \cdots a_n a_1 \cdots a_{j-1}$,
where the summation is over all indices $j$ such that $a_j = a$.
Then for a potential $\Phi \in \mathbb{C}\mathcal{G} / [\mathbb{C}\mathcal{G}, \mathbb{C}\mathcal{G}]$, which is some linear combination of cyclic paths in $\mathcal{G}$, we define the algebra
$A(\mathbb{C}\mathcal{G}, \Phi) = \mathbb{C}\mathcal{G} / \{ \partial_a \Phi \}$,
which is the quotient of the path algebra by the two-sided ideal generated by the relations $\partial_a \Phi \in \mathbb{C}\mathcal{G}$, for all edges $a$ of $\mathcal{G}$.
The Hilbert series $H_A$ for $A(\mathbb{C}\mathcal{G}, \Phi)$ is defined as $H_A(t) = \sum_{k=0}^{\infty} H_{ji}^k t^k$, where the $H_{ji}^k$ are matrices which count the dimension of the subspace $\{ i x j | \; x \in A(\mathbb{C}\mathcal{G}, \Phi)_k \}$, where $A(\mathbb{C}\mathcal{G}, \Phi)_k$ is the subspace of $A(\mathbb{C}\mathcal{G}, \Phi)$ of all paths of length $k$, and $i,j \in A(\mathbb{C}\mathcal{G}, \Phi)_0$.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=100mm]{fig-Oc-Kup.eps}\\
\caption{Cells associated to trivalent vertices} \label{fig:Oc-Kup}
\end{center}
\end{figure}
Ocneanu \cite{ocneanu:2000ii} defined a cell system $W$ on any $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$, associating a complex number $W \left( \triangle_{i,j,k}^{(a,b,c)} \right)$, now called an Ocneanu cell, to each closed loop of length three $\triangle_{i,j,k}^{(a,b,c)}$ in $\mathcal{G}$ as in Figure \ref{fig:Oc-Kup}, where $a,b,c$ are edges on $\mathcal{G}$, and $i,j,k$ are the vertices on $\mathcal{G}$ given by $i=s(a)=r(c)$, $j=s(b)=r(a)$, $k=s(c)=r(b)$.
These cells satisfy two properties, called Ocneanu's type I, II equations respectively, which are obtained by evaluating the Kuperberg relations K2, K3 for an $A_2$-spider \cite{kuperberg:1996} using the identification in Figure \ref{fig:Oc-Kup}: \\
$(i)$ for any type I frame \includegraphics[width=16mm]{fig_typeIframe.eps} in $\mathcal{G}$ we have
$$\sum_{k,b_1,b_2} W \left( \triangle_{i,j,k}^{(a,b_1,b_2)} \right) \overline{W \left( \triangle_{i,j,k}^{(a',b_1,b_2)} \right)} = \delta_{a,a'} [2]_q \phi_i \phi_j$$
$(ii)$ for any type II frame \includegraphics[width=30mm]{fig_typeIIframe.eps} in $\mathcal{G}$ we have
\begin{eqnarray*}
\lefteqn{ \sum_{k,b_j} \phi_k^{-1} W \left( \triangle_{i_2,i_1,k}^{(a_2,b_1,b_2)} \right) \overline{W \left( \triangle_{i_2,i_3,k}^{(a_3,b_3,b_2)} \right)} W \left( \triangle_{i_4,i_3,k}^{(a_4,b_3,b_4)} \right) \overline{W \left( \triangle_{i_4,i_1,k}^{(a_1,b_4,b_1)} \right)} } \nonumber \\
& \qquad & = \delta_{a_1,a_4} \delta_{a_2,a_3} \phi_{i_4} \phi_{i_1} \phi_{i_2} + \delta_{a_1,a_2} \delta_{a_3,a_4} \phi_{i_1} \phi_{i_2} \phi_{i_3} \hspace{25mm} \label{eqn:typeII_frame}
\end{eqnarray*}
Here $(\phi_v)_v$ is the Perron-Frobenius eigenvector for the Perron-Frobenius eigenvalue $\alpha = [3]_q$ of $\mathcal{G}$.
The existence of these cells for the finite $\mathcal{ADE}$ graphs was claimed by Ocneanu \cite{ocneanu:2000ii}, and shown in \cite{evans/pugh:2009i} with the exception of the graph $\mathcal{E}_4^{(12)}$. These cells define a unitary connection on the graph $\mathcal{G}$ which satisfy the Yang-Baxter equation \cite[Lemma 3.2]{evans/pugh:2009i}.
Two cell systems $W_1$, $W_2$ on an $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$ are equivalent if, for each pair of adjacent vertices $i$, $j$ of $\mathcal{G}$, we can find a family of unitary matrices $(u(a,b))_{a,b}$, where $a$, $b$ are any pair of edges from $i$ to $j$, such that
$$W_1(\triangle_{i_1,i_2,i_3}^{(a_1,a_2,a_3)}) = \sum_{a_1',a_2',a_3'} u(a_1,a_1') u(a_2,a_2') u(a_3,a_3') W_2(\triangle_{i_1,i_2,i_3}^{(a_1',a_2',a_3')}),$$
where $a_l$ are edges from $i_l$ to $i_{l+1}$, and the sum is over all edges $a_l'$ from $i_l$ to $i_{l+1}$, $l=1,2,3$.
There is up to equivalence precisely one connection on the graphs $\mathcal{A}^{(m)}$, $\mathcal{A}^{(2m+1)\ast}$, $\mathcal{E}^{(8)}$, $\mathcal{E}^{(8)\ast}$, $\mathcal{E}_5^{(12)}$ and $\mathcal{E}^{(24)}$. For the graphs $\mathcal{A}^{(2m)\ast}$ and $\mathcal{E}_2^{(12)}$ there are precisely two inequivalent connections, which are obtained from each other by a $\mathbb{Z}_2$ symmetry of the graph. This $\mathbb{Z}_2$ symmetry is the conjugation of the graph in the case of $\mathcal{E}_2^{(12)}$. There is at least one connection for each graph $\mathcal{D}^{(m)}$, $m \not \equiv 0 \textrm{ mod } 3$, and at least two inequivalent connections for each graph $\mathcal{D}^{(3p)}$, which are the complex conjugates of each other. There is at least one connection for each graph $\mathcal{D}^{(2m+1)\ast}$, and at least two inequivalent connections for each graph $\mathcal{D}^{(2m)\ast}$, which are obtained from each other by a $\mathbb{Z}_2$ symmetry of the graph. There are also at least two inequivalent connections for the graph $\mathcal{E}_1^{(12)}$, which are obtained from each other by conjugation of the graph.
For the $SU(3)$ $\mathcal{ADE}$ graphs, we define the almost Calabi-Yau algebra $A(\mathcal{G},W)$ to be the graded quotient algebra
$$A(\mathcal{G},W) := A(\mathbb{C} \mathcal{G}, \Phi_W),$$
where the potential $\Phi_W$ is given by \cite[equation (40)]{evans/pugh:2010ii} (see also \cite[Remark 4.5.7]{ginzburg:2006}):
$$\Phi_W = \sum_{abc} W(\triangle_{abc}) \triangle_{abc} \quad \in \mathbb{C} \mathcal{G} / [\mathbb{C} \mathcal{G}, \mathbb{C} \mathcal{G}],$$
where the summation is over all closed paths $abc$ of length 3 on $\mathcal{G}$. The grading on $\mathbb{C}\mathcal{G}$ descends to the quotient algebra $A = A(\mathcal{G},W)$.
These almost Calabi-Yau algebras were studied in \cite{evans/pugh:2010ii} for all the cell systems constructed in \cite{evans/pugh:2009i}.
Equivalent cell systems yield isomorphic almost Calabi-Yau algebras.
For any cell system $W$, we can take its complex conjugate $\overline{W}$ to obtain another (possibly equivalent) cell system. The almost Calabi-Yau algebra for $\overline{W}$ is isomorphic to that for $W$.
The conjugation $\tau: {}_N \mathcal{X}_N \rightarrow {}_N \mathcal{X}_N$ on the braided system of endomorphisms of $SU(3)_k$ on a factor $N$, given by the conjugation on the representations of $SU(3)$, induces a conjugation $\tau: {}_N \mathcal{X}_M \rightarrow {}_N \mathcal{X}_M$ such that $G_{\overline{\lambda}} = \tau G_{\lambda} \tau$, where $G_{\lambda} a = \lambda a$ for $\lambda \in {}_N \mathcal{X}_N$, $a \in {}_N \mathcal{X}_M$. For any cell system $W = W^+$, this conjugation of the graph yields a conjugate cell system $W^-$, which might be equivalent to $W^+$. The almost Calabi-Yau algebra for $W^-$ is anti-isomorphic to that for $W^+$.
The Hilbert series $H_A(t)$ of $A(\mathcal{G},W)$, for an $SU(3)$ $\mathcal{ADE}$ graph $\mathcal{G}$ with adjacency matrix $\Delta_{\mathcal{G}}$, Coxeter number $h=k+3$ and cell system $W$, is given by \cite[Theorem 3.1]{evans/pugh:2010ii}
\begin{equation} \label{eqn:Hilbert_Series-SU(3)ADE}
H_A (t) = \frac{1 - P t^h}{1 - \Delta_{\mathcal{G}} t + \Delta_{\mathcal{G}}^T t^2 - t^3},
\end{equation}
where $P$ is the permutation matrix corresponding to a $\mathbb{Z}_3$ symmetry of the graph.
It is the identity for $\mathcal{D}^{(n)}$, $\mathcal{A}^{(n)\ast}$, $n \geq 5$, $\mathcal{E}^{(8)\ast}$, $\mathcal{E}_l^{(12)}$, $l=1,2,4,5$, and $\mathcal{E}^{(24)}$. For the remaining graphs $\mathcal{A}^{(n)}$, $\mathcal{D}^{(n) \ast}$ and $\mathcal{E}^{(8)}$, let $V$ be the permutation matrix corresponding to the clockwise rotation of the graph by $2 \pi /3$. Then
$$ P = \left\{
\begin{array}{cl} V^2 & \mbox{ for } \quad \mathcal{A}^{(n)}, n \geq 4, \\
V & \mbox{ for } \quad \mathcal{E}^{(8)}, \\
V^{2n} & \mbox{ for } \quad \mathcal{D}^{(n) \ast}, n \geq 5.
\end{array} \right.$$
The numerator and denominator in (\ref{eqn:Hilbert_Series-SU(3)ADE}) commute, since any permutation matrix which corresponds to a symmetry of the graph $\mathcal{G}$ commutes with $\Delta_{\mathcal{G}}$ and $\Delta_{\mathcal{G}}^T$.
\subsection{Periodic resolution for almost Calabi-Yau algebras} \label{sect:resolution_almostCYalg}
We define a non-degenerate form on $A$ by setting $f$ to be the function which is 0 on every element of $A$ of length $< h-3$, and 1 on $u_{i\nu(i)}$ for some $i \in \mathcal{G}_1$, where $u_{j\nu(j)}$ denotes a generator of the one-dimensional top-degree space $j \cdot A_{h-3} \cdot \nu(j)$, where $\nu$ is the permutation of the vertices of $\mathcal{G}$ given by the permutation matrix $P$ in (\ref{eqn:Hilbert_Series-SU(3)ADE}). Then using the relation $(x,y) = (y,\beta(x))$ this determines the value of $f$ on $u_{j\nu(j)}$, for all other $j \in \mathcal{G}_1$. We normalize the $u_{j\nu(j)}$ such that $f(u_{j\nu(j)}) = 1$ for all $j \in \mathcal{G}_1$.
The image of the simple object $f_{(k,0)} \in A_2\textrm{-}TL^{(k)}$ under the functor $F$ given by (\ref{eqn:functorF}) defines a unique permutation $\nu$ of the graph $\mathcal{G}$, which is described as follows.
The permutation $\nu$ of the graph is given by the $\mathbb{Z}_3$ symmetry which defines the permutation matrix $P$ in (\ref{eqn:Hilbert_Series-SU(3)ADE}) (note that there are no double edges on the graphs $\mathcal{G}$ for which $P$ is non-trivial).
Then the Nakayama automorphism $\beta$ of $A$ is defined on $\mathcal{G}$ by $\beta = \nu$ \cite[Theorem 4.6]{evans/pugh:2010ii}.
Now $A$ has the following finite resolution as an $A$-$A$ bimodule \cite[Theorem 5.1]{evans/pugh:2010ii}:
\begin{equation} \label{exact_seq-almostCY}
0 \rightarrow \mathcal{N}[h] \stackrel{\iota_0}{\rightarrow} A \otimes_S A[3] \stackrel{\mu_3}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S A[1] \stackrel{\mu_2}{\rightarrow} A \otimes_S V \otimes_S A \stackrel{\mu_1}{\rightarrow} A \otimes_S A \stackrel{\mu_0}{\rightarrow} A \rightarrow 0.
\end{equation}
Here $S$ is the $A$-$A$ bimodule $(\mathbb{C}\mathcal{G})_0$, and $V$, $\widetilde{V}$ are the $A$-$A$ bimodules generated by $\mathcal{G}_1$, $\mathcal{G}^{\mathrm{op}}_1$ respectively. The $A$-$A$ bimodule $\mathcal{N} = {}_1 A_{\beta^{-1}}$ is equal to $A$ as a vector space. The left $A$-action is given by concatenation, but the right $A$-action is twisted by the inverse of the Nakayama automorphism $\beta$, i.e. $a \cdot x \cdot b = ax\beta^{-1}(b)$ for all $a,b \in A$, $x \in \mathcal{N}$.
The connecting $A$-$A$ bimodule maps are given by
\begin{eqnarray}
\mu_0(1 \otimes 1) & = & 1, \label{mu_0} \\
\mu_1(1 \otimes a \otimes 1) & = & a \otimes 1 - 1 \otimes a, \label{mu_1} \\
\mu_2(1 \otimes \widetilde{a} \otimes 1) & = & \sum_{b,b' \in \mathcal{G}_1} W_{abb'} (b \otimes b' \otimes 1 + 1 \otimes b \otimes b'), \label{mu_2} \\
\mu_3(1 \otimes 1) & = & \sum_{a \in \mathcal{G}_1} a \otimes \widetilde{a} \otimes 1 - \sum_{a \in \mathcal{G}_1} 1 \otimes \widetilde{a} \otimes a, \label{mu_3} \\
\iota_0(1) & = & \sum_j w_j \otimes w_j^{\ast}, \nonumber
\end{eqnarray}
where $\{ w_j \}$ is a homogeneous basis for $A$, and $\{ w_j^{\ast} \}$ is its corresponding dual basis, i.e. $w_j w_j^{\ast} = u_{i \nu(i)}$ where $i = s(w_j)$. The $A$-$A$ bimodule $B = B^{(1)} \otimes_S \cdots \otimes_S B^{(p)}$ is equipped with the \emph{total grading} which comes from the grading on the graded $A$-$A$ bimodules $B^{(i)}$, that is, $B = \bigoplus_{k=0}^{\infty} B_k$ where $B_k = \bigoplus_{k_i: \sum_{i=1}^p k_i = k} B^{(1)}_{k_1} \otimes_S \cdots \otimes_S B^{(p)}_{k_p}$.
For each $SU(3)$ $\mathcal{ADE}$ graph, the Nakayama automorphism has order 3, $\beta^{3} = \mathrm{id}$, so we can make a canonical identification $A = \mathcal{N} \otimes_{A} \mathcal{N} \otimes_{A} \mathcal{N}$. We let $\mathcal{N}^{(k)} := {}_1 A_{\beta^{-k}}$, for $k \in \mathbb{Z}$. In particular, we have $A = \mathcal{N}^{(0)}$, $\mathcal{N} = \mathcal{N}^{(1)}$ and $\mathcal{N}^{(2)} = {}_1 A_{\beta} = \mathcal{N} \otimes_{A} \mathcal{N}$. Note that for graphs with trivial Nakayama automorphism, $A = \mathcal{N}^{(k)}$ as $A$-$A$ bimodules, for all $k \in \mathbb{Z}$.
Applying the functor $- \otimes_{A} \mathcal{N}$ to the exact sequence (\ref{exact_seq-almostCY}) we obtain the exact sequence:
\begin{eqnarray*}
& 0 \rightarrow \mathcal{N}^{(2)}[2h] \stackrel{\iota_2}{\rightarrow} A \otimes_S \mathcal{N}[h+3] \stackrel{\mu_7}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S \mathcal{N}[h+1] \stackrel{\mu_6}{\rightarrow} A \otimes_S V \otimes_S \mathcal{N}[h] & \nonumber \\
& \stackrel{\mu_5}{\rightarrow} A \otimes_S \mathcal{N}[h] \stackrel{\iota_1}{\rightarrow} \mathcal{N}[h] \rightarrow 0, & \label{exact_seq-almostCY-2}
\end{eqnarray*}
where $\iota_1(x \otimes y) = xy$, $\iota_2(a) = a \sum_j w_j \otimes w_j^{\ast}$, where $\{ w_j \}$ is a homogeneous basis for $A$ and $\{ w_j^{\ast} \}$ is its corresponding dual basis, and $\mu_{i+4} = \mu_i$. Similarly, applying the functor a second time we obtain the exact sequence:
\begin{eqnarray*}
& 0 \rightarrow A[3h] \stackrel{\iota_4}{\rightarrow} A \otimes_S \mathcal{N}^{(2)}[2h+3] \stackrel{\mu_{11}}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S \mathcal{N}^{(2)}[2h+1] \stackrel{\mu_{10}}{\rightarrow} A \otimes_S V \otimes_S \mathcal{N}^{(2)}[2h] & \nonumber \\
& \stackrel{\mu_9}{\rightarrow} A \otimes_S \mathcal{N}^{(2)}[2h] \stackrel{\iota_3}{\rightarrow} \mathcal{N}^{(2)}[2h] \rightarrow 0. & \label{exact_seq-almostCY-3}
\end{eqnarray*}
We now construct a projective resolution of $A$, that is, a resolution of $A$ by projective modules.
Setting $\mu_4 = \iota_0 \iota_1$, $\mu_8 = \iota_2 \iota_3$, we obtain the following projective resolution of $A$, which is periodic with period 12:
\begin{eqnarray}
& \cdots \,\, \rightarrow A[3h] \stackrel{\mu_{12}}{\rightarrow} A \otimes_S \mathcal{N}^{(2)}[2h+3] \stackrel{\mu_{11}}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S \mathcal{N}^{(2)}[2h+1] \stackrel{\mu_{10}}{\rightarrow} A \otimes_S V \otimes_S \mathcal{N}^{(2)}[2h] & \nonumber \\
& \stackrel{\mu_9}{\rightarrow} A \otimes_S \mathcal{N}^{(2)}[2h] \stackrel{\mu_8}{\rightarrow} A \otimes_S \mathcal{N}[h+3] \stackrel{\mu_7}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S \mathcal{N}[h+1] \stackrel{\mu_6}{\rightarrow} A \otimes_S V \otimes_S \mathcal{N}[h] & \nonumber \\
& \stackrel{\mu_5}{\rightarrow} A \otimes_S \mathcal{N}[h] \stackrel{\mu_4}{\rightarrow} A \otimes_S A[3] \stackrel{\mu_3}{\rightarrow} A \otimes_S \widetilde{V} \otimes_S A[1] \stackrel{\mu_2}{\rightarrow} A \otimes_S V \otimes_S A \stackrel{\mu_1}{\rightarrow} A \otimes_S A \stackrel{\mu_0}{\rightarrow} A \rightarrow 0, & \nonumber \\
&& \label{resolution-almostCY}
\end{eqnarray}
where the connecting maps $\mu_i$ are given by (\ref{mu_0})-(\ref{mu_3}) for $0 \leq i \leq 3$, $\mu_4(x \otimes y) = xy \sum_j w_j \otimes w_j^{\ast}$, where $\{ w_j \}$ is a homogeneous basis for $A$ and $\{ w_j^{\ast} \}$ is its corresponding dual basis, and $\mu_i = \mu_{i-4}$ for $i \geq 5$.
Thus we find that the Hochschild (co)homology of $A$ is periodic with period 12, i.e. the grading is shifted by $3h$ ($-3h$) when the degree of the homology (respectively cohomology) is shifted by 12.
In the case of trivial Nakayama automorphism the Hochschild (co)homology of $A$ in fact has period 4.
\section{The Hochschild homology of $A(\mathcal{G},W)$} \label{sect:Hoch_hom}
\subsection{The Hochschild homology complex} \label{sect:Hoch_hom-complex}
In this section we will construct a complex which determines the Hochschild homology of the almost Calabi-Yau algebra $A = A(\mathcal{G},W)$.
Let $A^{\mathrm{op}}$ denote the algebra with opposite multiplication, i.e. $a \cdot b = ba$, and define $A^e = A^{\mathrm{op}} \otimes_S A$. Any $A$-$A$ bimodule becomes a left $A^e$-module, and vice versa, by defining the left action of $A^e$ on $A$ by $(a \otimes b) x = bxa$ for all $x \in A$, $a \otimes b \in A^{\mathrm{op}} \otimes_S A$.
The Hochschild homology $HH_{\bullet}(A)$ of $A$ may be defined to be the derived functor $HH_n(A) = \mathrm{Tor}_n^{A^e}(A,A)$, e.g. \cite[Proposition 1.1.13]{loday:1998}, i.e. as the homology of the complex
$$\cdots \rightarrow P_2 \otimes_{A^e} A \rightarrow P_1 \otimes_{A^e} A \rightarrow P_0 \otimes_{A^e} A \rightarrow A \otimes_{A^e} A \rightarrow 0$$
where $\quad \cdots \rightarrow P_2 \rightarrow P_1 \rightarrow P_0 \rightarrow A \rightarrow 0$ is any projective resolution of $A$.
For an $A$-$A$ bimodule $M$, denote by $M^S$ the $S$-centralizer sub-bimodule given by all elements $x \in M$ such that $i x = x i$ for all $i \in S$.
We make the following identifications, for $k=0,1,2$ (c.f. \cite{etingof/eu:2007}):
$$\begin{array}{ll}
(A \otimes_S \mathcal{N}^{(k)}) \otimes_{A^e} A = (\mathcal{N}^{(k)})^S: & (x \otimes y) \otimes z = y\beta^{-k}(zx), \\
(A \otimes_S V \otimes_S \mathcal{N}^{(k)}) \otimes_{A^e} A = (V \otimes_S \mathcal{N}^{(k)})^S: & (x \otimes a \otimes y) \otimes z = a \otimes y\beta^{-k}(zx), \\
(A \otimes_S \widetilde{V} \otimes_S \mathcal{N}^{(k)}) \otimes_{A^e} A = (\widetilde{V} \otimes_S \mathcal{N}^{(k)})^S: & (x \otimes \widetilde{a} \otimes y) \otimes z = \widetilde{a} \otimes y\beta^{-k}(zx),
\end{array}$$
where the left and right hand sides have the same total degree.
Thus, applying the functor $- \otimes_{A^e} A$ to the resolution (\ref{resolution-almostCY}), we obtain the Hochschild homology complex:
\begin{eqnarray}
& \cdots \,\, \rightarrow A^S[3h] \stackrel{\mu_{12}'}{\rightarrow} (\mathcal{N}^{(2)})^S[2h+3] \stackrel{\mu_{11}'}{\rightarrow} (\widetilde{V} \otimes_S \mathcal{N}^{(2)})^S[2h+1] \stackrel{\mu_{10}'}{\rightarrow} (V \otimes_S \mathcal{N}^{(2)})^S[2h] & \nonumber \\
& \stackrel{\mu_9'}{\rightarrow} (\mathcal{N}^{(2)})^S[2h] \stackrel{\mu_8'}{\rightarrow} \mathcal{N}^S[h+3] \stackrel{\mu_7'}{\rightarrow} (\widetilde{V} \otimes_S \mathcal{N})^S[h+1] \stackrel{\mu_6'}{\rightarrow} (V \otimes_S \mathcal{N})^S[h] & \nonumber \\
& \stackrel{\mu_5'}{\rightarrow} \mathcal{N}^S[h] \stackrel{\mu_4'}{\rightarrow} A^S[3] \stackrel{\mu_3'}{\rightarrow} (\widetilde{V} \otimes_S A)^S[1] \stackrel{\mu_2'}{\rightarrow} (V \otimes_S A)^S \stackrel{\mu_1'}{\rightarrow} A^S \rightarrow 0, & \label{HH_hom_complex-almostCY}
\end{eqnarray}
where the connecting maps are given, for $k=0,1,2,\ldots \;\;$ by
\begin{eqnarray*}
\mu_{4k+1}'(a \otimes x) & = & \mu_{4k+1}(1 \otimes a \otimes 1) \otimes_{A^e} \beta^{k}(x) = (a \otimes 1 - 1 \otimes a) \otimes_{A^e} \beta^{k}(x) \\
& = & x \beta^{-k}(a) - ax, \\
\mu_{4k+2}'(\widetilde{a} \otimes x) & = & \mu_{4k+2}(1 \otimes \widetilde{a} \otimes 1) \otimes_{A^e} \beta^{k}(x) \\
& = & \sum_{b,b' \in \mathcal{G}_1} W_{abb'} (b \otimes b' \otimes 1 + 1 \otimes b \otimes b') \otimes_{A^e} \beta^{k}(x) \\
& = & \sum_{b,b' \in \mathcal{G}_1} W_{abb'} (b' \otimes x \beta^{-k}(b) + b \otimes b'x), \\
\mu_{4k+3}'(x) & = & \mu_{4k+3}(1 \otimes 1) \otimes_{A^e} \beta^{k}(x) = \left( \sum_{a \in \mathcal{G}_1} a \otimes \widetilde{a} \otimes 1 - 1 \otimes \widetilde{a} \otimes a \right) \otimes_{A^e} \beta^{k}(x) \\
& = & \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes (x \beta^{-k}(a) - ax), \\
\mu_{4k+4}'(y) & = & \mu_{4k+4}(1 \otimes 1) \otimes_{A^e} \beta^{k+1}(y) = \left( \sum_j w_j \otimes w_j^{\ast} \right) \otimes_{A^e} \beta^{k+1}(y) \\
& = & \sum_j w_j^{\ast} \beta(y) \beta^{-k}(w_j),
\end{eqnarray*}
where $a \in V$, $x \in \mathcal{N}^{(k)}$, $y \in \mathcal{N}^{(k+1)}$, $\{ w_j \}$ is a homogeneous basis for $A$ and $\{ w_j^{\ast} \}$ is its corresponding dual basis.
We will now show that this complex has a self-duality.
Using the non-degenerate form, we can make the identifications $\mathcal{N}^{(k)} = (\mathcal{N}^{(2-k)})^{\ast}[h-3]$ by sending $x \mapsto (-,x)$.
We can define a non-degenerate form on $(V \oplus \widetilde{V}) \otimes_S \mathcal{N}^{(k)}$ by $(a_1 \otimes x_1, a_2 \otimes x_2) = \delta_{a_1,\beta^{k-1}(\widetilde{a_2})} (x_1,x_2)$ for $x_1 \in \mathcal{N}^{(2-k)}$, $x_2 \in \mathcal{N}^{(k)}$, and $a_1 \in V_1$, $a_2 \in V_2$, where $V_i \in \{ V, \widetilde{V}\}$, $i=1,2$. For the $\mathcal{A}^{\ast}$ graphs, $V = \widetilde{V}$ and we replace $(V \oplus \widetilde{V}) \otimes_S \mathcal{N}^{(k)}$ above by $V \otimes_S \mathcal{N}^{(k)}$.
This allows us to make identifications $V \otimes_S \mathcal{N}^{(k)} = (\widetilde{V} \otimes_S \mathcal{N}^{(2-k)})^{\ast}[h-1]$, $\widetilde{V} \otimes_S \mathcal{N}^{(k)} = (V \otimes_S \mathcal{N}^{(2-k)})^{\ast}[h-1]$, by sending $a \otimes x \mapsto (-,a \otimes x)$.
If we take the Hochschild homology sequence (\ref{HH_hom_complex-almostCY}) and dualise, we get:
\begin{eqnarray*}
& \cdots \,\, \stackrel{(\mu_{12}')^{\ast}}{\leftarrow} A^S[-3h] \stackrel{(\mu_{11}')^{\ast}}{\leftarrow} (V \otimes_S A)^S[-3h] \stackrel{(\mu_{10}')^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S A)^S[-3h+1] \stackrel{(\mu_9')^{\ast}}{\leftarrow} & \\
& \stackrel{(\mu_9')^{\ast}}{\leftarrow} A^S[-3h+3] \stackrel{(\mu_8')^{\ast}}{\leftarrow} \mathcal{N}^S[-2h] \stackrel{(\mu_7')^{\ast}}{\leftarrow} (V \otimes_S \mathcal{N})^S[-2h] \stackrel{(\mu_6')^{\ast}}{\leftarrow} & \\
& \stackrel{(\mu_6')^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S \mathcal{N})^S[-2h+1] \stackrel{(\mu_5')^{\ast}}{\leftarrow} \mathcal{N}^S[-2h+3] \stackrel{(\mu_4')^{\ast}}{\leftarrow} (\mathcal{N}^{(2)})^S[-h] \stackrel{(\mu_3')^{\ast}}{\leftarrow} & \\
& \stackrel{(\mu_3')^{\ast}}{\leftarrow} (V \otimes_S \mathcal{N}^{(2)})^S[-h] \stackrel{(\mu_2')^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S \mathcal{N}^{(2)})^S[-h+1] \stackrel{(\mu_1')^{\ast}}{\leftarrow} (\mathcal{N}^{(2)})^S[-h+3] \leftarrow 0. &
\end{eqnarray*}
\begin{Prop}
We have $\mu_i' = \pm(\mu_{12-i}')^{\ast}$, $i=1,\ldots,11$.
\end{Prop}
\noindent \emph{Proof}:
(i) $\mu_1' = - (\mu_{11}')^{\ast}$:
Let $a \in V$, $x \in A$ and $y \in \mathcal{N}^{(2)}$. Then
\begin{eqnarray*}
(\mu_1'(a \otimes x),y) & = & (xa-ax,y) = (x,ay-y\beta(a)) = (a \otimes x, -\sum_{b \in \mathcal{G}_1} \widetilde{b} \otimes (y\beta(b)-by)) \\
& = & (a \otimes x,-\mu_{11}'(y)).
\end{eqnarray*}
(ii) $\mu_2' = (\mu_{10}')^{\ast}$:
Let $a,a' \in V$, $x \in A$ and $y \in \mathcal{N}^{(2)}$. Then
\begin{eqnarray*}
\lefteqn{ \hspace{-10mm} (\mu_2'(\widetilde{a} \otimes x),\widetilde{a'} \otimes y) \;\; = \;\; (\sum_{b,b' \in \mathcal{G}_1} W_{abb'} (b' \otimes xb + b \otimes b'x), \widetilde{a'} \otimes y) } \\
\hspace{10mm} & = & (\sum_{b \in \mathcal{G}_1} W_{aba'} xb + \sum_{b' \in \mathcal{G}_1} W_{aa'b'} b'x,y) \;\; = \;\; (x, \sum_{b \in \mathcal{G}_1} W_{aba'} by + \sum_{b' \in \mathcal{G}_1} W_{aa'b'} y\beta(b')) \\
& = & (\widetilde{a} \otimes x, \sum_{b,b' \in \mathcal{G}_1} W_{b'ba'} (b' \otimes by + b \otimes y\beta(b')) = (\widetilde{a} \otimes x, \mu_{10}'(\widetilde{a'} \otimes y)).
\end{eqnarray*}
(iii) $\mu_3' = - (\mu_9')^{\ast}$:
Let $a' \in V$, $x \in A$ and $y \in \mathcal{N}^{(2)}$. Then
\begin{eqnarray*}
(\mu_3'(x),a' \otimes y) & = & (\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes (xa-ax),a' \otimes y) = (xa'-a'x,y) = (x,a'y-y\beta(a')) \\
& = & (x,-\mu_9'(a' \otimes y)).
\end{eqnarray*}
(iv) $\mu_4' = (\mu_8')^{\ast}$:
Let $x \in \mathcal{N}$ and $y \in \mathcal{N}^{(2)}$. Then
\begin{eqnarray*}
(\mu_4'(x),y) & = & (\sum_j w_j^{\ast}\beta(x)w_j,y) = (\sum_j w_j x \beta^2(w_j^{\ast}),y) = (x,\sum_j \beta^2(w_j^{\ast})y\beta(w_j)) \\
& = & (x,\sum_j w_j^{\ast} \beta(y) \beta^2(w_j)) = (x,\mu_8'(y)),
\end{eqnarray*}
where the second equality holds since if $\{ w_j^{\ast} \}$ is a dual basis of $\{ w_j \}$, then $\{ w_j \}$ is a dual basis of $\{ \beta^2(w_j^{\ast}) \}$, and $\sum_j w_j^{\ast}\beta(x)w_j = 0 = \sum_j w_j x \beta^2(w_j^{\ast})$ unless $|x|=0$ such that $\beta(x)=x$. The penultimate equality is given by replacing the basis $\{ \beta^2(w_j^{\ast}) \}$ with the equivalent basis $\{ w_j^{\ast} \}$, and the fact that $\sum_j \beta^2(w_j^{\ast})y\beta(w_j) = 0 = \sum_j w_j^{\ast} \beta(y) \beta^2(w_j)$ unless $|y|=0$ such that $\beta(y)=y$. \\
(v) $\mu_5' = - (\mu_7')^{\ast}$:
Let $a \in V$ and $x, y \in \mathcal{N}$. Then
\begin{eqnarray*}
(\mu_5'(a \otimes x),y) & = & (x\beta^2(a)-ax,y) = (x,\beta^2(a)y-y\beta(a)) \\
& = & (a \otimes x, -\sum_{b \in \mathcal{G}_1} \widetilde{b} \otimes (y\beta^2(b)-by)) = (a \otimes x,-\mu_7'(y)).
\end{eqnarray*}
(vi) $\mu_6' = (\mu_6')^{\ast}$:
Let $a,a' \in V$ and $x, y \in \mathcal{N}$. Then
\begin{eqnarray*}
\lefteqn{ \hspace{-2mm} (\mu_6'(\widetilde{a} \otimes x),\widetilde{a'} \otimes y) \;\; = \;\; (\sum_{b,b' \in \mathcal{G}_1} W_{abb'} (b' \otimes x\beta^2(b) + b \otimes b'x), \widetilde{a'} \otimes y) } \\
& = & (\sum_{b \in \mathcal{G}_1} W_{aba'} x\beta^2(b) + \sum_{b' \in \mathcal{G}_1} W_{aa'b'} b'x,y) \;\; = \;\; (x, \sum_{b \in \mathcal{G}_1} W_{aba'} \beta^2(b)y + \sum_{b' \in \mathcal{G}_1} W_{aa'b'} y\beta(b')) \\
& = & (\widetilde{a} \otimes x, \sum_{b,b' \in \mathcal{G}_1} W_{b'ba'} (b' \otimes by + b \otimes y\beta^2(b')) = (\widetilde{a} \otimes x, \mu_6'(\widetilde{a'} \otimes y)).
\end{eqnarray*}
\vspace{-5mm} $\,$
\hfill
$\Box$
Note however that $(\mu_{12}')^{\ast} = \mu_{12}' \circ \beta$:
Let $x,y \in A$. Then
\begin{eqnarray*}
(\mu_{12}'(x),y) & = & (\sum_j w_j^{\ast}\beta(x)\beta(w_j),y) = (\beta(x),\sum_j \beta(w_j) y \beta(w_j^{\ast})) = (x,\sum_j w_j\beta^2(y)w_j^{\ast}) \\
& = & (x,\sum_j w_j^{\ast}\beta^2(y)\beta(w_j)) = (x,\mu_{12}'(\beta(y))),
\end{eqnarray*}
where the penultimate equality holds since if $\{ w_j^{\ast} \}$ is a dual basis of $\{ w_j \}$, then $\{ \beta(w_j) \}$ is a dual basis of $\{ w_j^{\ast} \}$.
From the self-duality of the Hochschild homology complex (\ref{HH_hom_complex-almostCY}) and $(\mu_{12}')^{\ast} = \mu_{12}' \circ \beta$, we have
\begin{eqnarray*}
HH_i(A)^{\ast} & \cong & HH_{11-i}(A)[3h], \qquad \qquad \; i=1,\ldots,10, \\
HH_{11}(A)^{\ast} & \cong & HH_{12}(A)[6h].
\end{eqnarray*}
The reduced Hochschild homology $\overline{HH}_{\bullet}(A)$ is defined as $\overline{HH}_0(A) = HH_0(A)/S$ and $\overline{HH}_n(A) = HH_n(A)$, $n>0$.
\subsection{The cyclic homology of $A(\mathcal{G},W)$} \label{sect:cyclic_hom}
Before we determine the Hochschild homology of $A(\mathcal{G},W)$ for certain $SU(3)$ $\mathcal{ADE}$ graphs, we introduce cyclic homology.
We begin by introducing the differential graded algebra $\Omega^{\bullet}A$ of non-commutative forms of $A$, and the non-commutative de Rham homology.
The $A$-$A$ bimodule $\Omega^1 A$ of non-commutative relative 1-forms on $A$ is defined as the kernel of the multiplication map $A \otimes_S A \rightarrow A$. The differential graded algebra $\Omega^{\bullet}A$ of non-commutative forms of $A$ is obtained by taking tensor powers of $\Omega^1 A$. The graded commutator in $\Omega^{\bullet}A$ is given by $[\omega,\omega'] = \omega \omega' - (-1)^{|\omega| |\omega'|} \omega' \omega$, where $|\omega| = n$ denotes the homological degree of $\omega \in \Omega^n A$. The reduced non-commutative de Rham homology of $A$ is defined by
$$\overline{H}DR_n (A) := H_n (\Omega^{\bullet}A/(S+[\Omega^{\bullet}A,\Omega^{\bullet}A]),d),$$
where the natural differential $\Omega^{\bullet}A \rightarrow \Omega^{\bullet+1}A$ descends to a de Rham differential on $\Omega^{\bullet}A/(S+[\Omega^{\bullet}A,\Omega^{\bullet}A])$.
Since $A$ is an augmented $S$-algebra, i.e. $A_0 = S$ and there is an augmentation $\varphi:A \rightarrow S$ such that $\varphi(1)=1$, by the non-commutative Poincar\'{e} lemma \cite{kontsevich:1993} (see also \cite[Lemma 4.5]{mejias:2002}), $\overline{H}DR_n (A) = \overline{H}DR_n (S) = 0$ for all $n$. Thus, from \cite[Lemma 3.6.1]{etingof/ginzburg:2007}, there is an exact sequence
\begin{equation} \label{seq:Connes_exact_seq}
0 \longrightarrow \overline{HH}_0(A) \stackrel{B}{\longrightarrow} \overline{HH}_1(A) \stackrel{B}{\longrightarrow} \overline{HH}_2(A) \stackrel{B}{\longrightarrow} \overline{HH}_0(A) \longrightarrow \cdots
\end{equation}
where $B$ is the Connes differential, which is degree-preserving, and the reduced cyclic homology of $A$ can be defined by
$$\overline{HC}_n (A) = \mathrm{ker}(B: \overline{HH}_{n+1}(A) \rightarrow \overline{HH}_{n+2}(A)) = \mathrm{Im}(B: \overline{HH}_{n}(A) \rightarrow \overline{HH}_{n+1}(A)).$$
The usual cyclic homology is related to the reduced cyclic homology by $\overline{HC}_0(A) = HC_0(A)/S$ and $\overline{HC}_n(A) = HC_n(A)$, $n>0$.
The (graded) Euler characteristic of the reduced cyclic homology is the polynomial in $t$ defined by $\chi_{\overline{HC}(A)}(t) = \sum_{i=0}^{\infty} (-1)^i H_{\overline{HC}_i(A)}(t)$.
It turns out to be easier to describe the Euler characteristic of $\mathrm{Sym}(\overline{HC}(A))_+$, where if $\chi_{\overline{HC}(A)}(t) = \sum_{k=0}^{\infty} a_k t^k$ then $\chi_{\mathrm{Sym}(\overline{HC}(A))_+}(t) = \prod_{k=1}^{\infty} (1-t^k)^{a_k}$.
In \cite[Prop. 3.7.1]{etingof/ginzburg:2007} it was shown that for $A$ the preprojective algebra of a non-Dynkin quiver,
\begin{equation} \label{eqn:Euler=prod_det}
\prod_{k=1}^{\infty} (1-t^k)^{-a_k} = \prod_{s=1}^{\infty} \mathrm{det}H_A(t^s),
\end{equation}
where $H_A(t)$ is the Hilbert series of $A$. The result (\ref{eqn:Euler=prod_det}) was extended to the case where $A$ is a Calabi-Yau algebra of dimension 3 in \cite[Prop. 5.4.9]{ginzburg:2006}.
In the case when $A$ is the almost Calabi-Yau algebra $A=A(\mathcal{G},W)$, the differential graded algebra $\mathfrak{D}_{\bullet} = T_S(V \oplus V^{\ast} \oplus S^{\ast})$ in \cite[Prop. 5.4.9]{ginzburg:2006} is no longer exact. However, we can build a larger free differential graded algebra $\mathfrak{D}_{\bullet}'$ by adding generators $x_n \in \mathfrak{D}_n'$ whose images under the differential give a basis for $H_n(\mathfrak{D}_{\bullet}')$, for each $n>0$. These generators lie in degree $nh$, where $h$ is the Coxeter number of $\mathcal{G}$. Then $\mathfrak{D}_{\bullet}'$ gives a free resolution of $A$, and a correction term corresponding to the numerator $1-Pt^{hs}$ of $H_A(t^s)$ appears in the formula (\ref{eqn:Euler=prod_det}). Thus the result (\ref{eqn:Euler=prod_det}) holds for the almost Calabi-Yau algebra $A=A(\mathcal{G},W)$ (c.f. \cite[Lemma 4.4.1]{etingof/eu:2007} in the case where $A$ is the preprojective algebra of a Dynkin quiver).
\subsection{$HH_0(A)$ for $A=A(\mathcal{G},W)$} \label{sect:HH_0}
In this section we compute the zeroth Hochschild homology $HH_0(A) = \mathrm{ker}(\mu_0')/\mathrm{Im}(\mu_1') = A/[A,A]$ for the simplest graphs, namely the graphs $\mathcal{A}^{(n)}$, $n \geq 4$, $\mathcal{D}^{(3k+3)}$, $k \geq 1$, $\mathcal{A}^{(n)\ast}$, $n \geq 5$, $\mathcal{D}^{(n)\ast}$, $n \geq 5$, $\mathcal{E}^{(8)}$ and $\mathcal{E}^{(8)\ast}$.
For a graded algebra $B = \bigoplus_{k=0}^{\infty} B_k$, let $B_+$ denote the positive degree part $B_+ = \bigoplus_{k=1}^{\infty} B_k$. For any $a,b \in A_+$ such that $r(a) = s(b)$ and $s(a) \neq r(b)$, $[a,b] = ab$, thus any non-cyclic path $ab$ is in $[A,A]$.
For $a,b \in A_+$ such that $r(a) = s(b)$ and $s(a) = r(b)$, $[a,b] = ab-ba$, thus cyclic paths are equivalent in $A/[A,A]$ if one is a cyclic permutation of the other.
Thus to determine $A/[A,A]$ we first consider all cyclic paths in $i A i$ for some $i \in \mathcal{G}_0$, then consider all cyclic paths in $j A j$ which do not pass through the vertex $i \neq j$, for some $j \in \mathcal{G}_0$, and so on.
Note that $S \hookrightarrow A/[A,A]$ since $[i,j]=0$ for all $i,j \in S$ and $[a,b] \subset A_+$ if either $a$ or $b$ have non-zero length.
\subsubsection{The identity $\mathcal{A}^{(n)}$ graphs}
The unique cell system $W$ (up to equivalence) was computed in \cite[Theorem 5.1]{evans/pugh:2009i}.
For the graph $\mathcal{A}^{(n)}$, $n \geq 4$, the space of cyclic paths $(0,0) A_+ (0,0) = 0$. Thus for any vertex $i \neq (0,0)$, any cyclic path $x \in i A_+ i$ which passes through $(0,0)$ is a cyclic permutation of a cyclic path $x' \in (0,0) A_+ (0,0)$.
Similarly, any cyclic path $x \in i A_+ i$ which does not pass through $(0,0)$ can be transformed by a combination the relations in $A$ and cyclic permutations to a cyclic path $x' \in (0,0) A_+ (0,0)$. Thus any cyclic path $x \in i A_+ i$ will be zero in $A/[A,A]$,
and we obtain
\begin{equation}
HH_0(A) \cong S.
\end{equation}
\subsubsection{The orbifold $\mathcal{D}^{(3k+3)}$ graphs}
We now consider the graphs $\mathcal{D}^{(3k+3)}$, $k \geq 1$, which are $\mathbb{Z}_3$-orbifolds of $\mathcal{A}^{(3k+3)}$. The graph $\mathcal{D}^{(9)}$ is illustrated in Figure \ref{fig:D(9)}. The weights $W(\triangle)$ for $\mathcal{A}^{(3k+3)}$ are invariant under the $\mathbb{Z}_3$ symmetry of the graph given by rotation by $2\pi/3$. Thus there is an orbifold solution for the cell system $W$ on $\mathcal{D}^{(3k+3)}$ where the weights $W(\triangle)$ are given by the corresponding weights for $\mathcal{A}^{(3k+3)}$ \cite[Theorem 6.2]{evans/pugh:2009i}. More precisely, excluding triangles $\triangle$ which contain one of the triplicated vertices $(k,k)_l$, the weight $W(\triangle_{i_1,i_2,i_3})$ for the triangle $\triangle_{i_1,i_2,i_3} = i_1 \rightarrow i_2 \rightarrow i_3 \rightarrow i_1$ on $\mathcal{D}^{(3k+3)}$ is given by the weight $W(\triangle_{i_1^{(0)},i_2^{(1)},i_3^{(2)}}) = W(\triangle_{i_1^{(1)},i_2^{(2)},i_3^{(0)}}) = W(\triangle_{i_1^{(2)},i_2^{(0)},i_3^{(1)}})$ for $\mathcal{A}^{(3k+3)}$, where $i_l^{(0)}$, $i_l^{(1)}$, $i_l^{(2)}$ are the three vertices of $\mathcal{A}^{(3k+3)}$ which are identified under the $\mathbb{Z}_3$ action to give the vertex $i_l$ of $\mathcal{D}^{(3k+3)}$, $l=1,2,3$.
If for a triangle $\triangle_{i_1,i_2,i_3}$ on $\mathcal{D}^{(3k+3)}$ there is no choice of vertices $i_1^{(j_1)}$, $i_2^{(j_2)}$, $i_3^{(j_3)}$ on $\mathcal{A}^{(3k+3)}$ which lie on a closed loop of length three $i_1^{(j_1)} \rightarrow i_2^{(j_2)} \rightarrow i_3^{(j_3)} \rightarrow i_1^{(j_1)}$, then we have $W(\triangle_{i_1,i_2,i_3}) = 0$.
The weight $W(\triangle)$ for a triangle $\triangle$ which contain one of the triplicated vertices $(k,k)_l$ is just given by one third of the weight for the corresponding triangle on $\mathcal{A}^{(3k+3)}$.
Thus the relations for $\mathcal{D}^{(3k+3)}$ are given precisely by the relations for $\mathcal{A}^{(3k+3)}$, except for the relations $\rho_{\gamma}$, $\rho_{\gamma'}$, which involve the triplicated vertices $(k,k)_l$.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=55mm]{fig-D9}\\
\caption{Graph $\mathcal{D}^{(9)}$} \label{fig:D(9)}
\end{center}
\end{figure}
Any cyclic path on $\mathcal{A}^{(3k+3)}$ yields a cyclic paths on $\mathcal{D}^{(3k+3)}$ by the above orbifold procedure. These cyclic paths will be zero in $A/[A,A]$, except for those which pass along the double edge of $\mathcal{D}^{(3k+3)}$ -- although these paths can be made to pass through $(0,0)$ in $A'/[A',A']$ for $A'=A(\mathcal{A}^{(3k+3)},W)$, when we do this for $A=A(\mathcal{D}^{(3k+3)},W)$ we obtain a cyclic path which passes through the vertex 1 of $\mathcal{D}^{(3k+3)}$ which corresponds to $(0,0)$ on $\mathcal{A}^{(3k+3)}$, but also a linear combination of cyclic paths which do not pass through the vertex 1, due to the fact that relations involving the double edge are not of the form $x=\lambda x'$ for basis paths $x,x' \in A$.
There are also cyclic paths in $A$ which do not come from cyclic paths in $A'$ by the orbifold procedure. These paths must necessarily pass along the double edge $(\gamma,\gamma')$ of $\mathcal{D}^{(3k+3)}$. Using the relations in $A$ and cyclic permutations, we can transform any such cyclic path, necessarily of length $3j$, $j \in \mathbb{N}$, due to the three-colourability of $\mathcal{D}^{(3k+3)}$, to a linear combination of cyclic basis paths $[(i_1 i_2 k_l i_1)^j]$, $l=1,2,3$, where $i_1 = s(\gamma)$, $i_2 = r(\gamma)$, $k_l := (k,k)_l$, and $x^m$ denotes the path $x x \cdots x$ ($m$ times). These basis paths are not equivalent in $[A,A]$, except when $j=k$ where $[(i_1 i_2 k_1 i_1)^{k}] = [(i_1 i_2 k_2 i_1)^{k}] = [(i_1 i_2 k_3 i_1)^{k}]$, thus
\begin{equation}
HH_0(A) \cong S \oplus C,
\end{equation}
where the graded vector space $C = \bigoplus_{j=1}^{k-1} \bigoplus_{l=1}^3 \mathbb{C} [(i_1 i_2 k_l i_1)^j] \oplus \mathbb{C} [(i_1 i_2 k_1 i_1)^{k}]$, and has Hilbert series $H_C(t) = \sum_{j=1}^{k-1} 3t^{3j} + t^{3k}$.
\subsubsection{The conjugate $\mathcal{A}^{\ast}$ graphs}
\begin{figure}[tb]
\begin{minipage}[t]{6cm}
\begin{center}
\includegraphics[width=35mm]{fig-A7,8star}\\
\caption{Graphs $\mathcal{A}^{(7)\ast}$, $\mathcal{A}^{(8)\ast}$} \label{fig:A(7,8)star}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{9cm}
\begin{center}
\includegraphics[width=90mm]{fig-D7,8star}\\
\caption{Graphs $\mathcal{D}^{(7)\ast}$, $\mathcal{D}^{(8)\ast}$} \label{fig:D(7,8)star}
\end{center}
\end{minipage}
\end{figure}
The unique cell system $W$ (up to equivalence) was computed in \cite[Theorems 7.1, 7.3 \& 7.4]{evans/pugh:2009i}, and we use the same notation for the cells here.
The $\mathcal{A}^{(n)\ast}$ graphs are illustrated in \cite[Figure 11]{evans/pugh:2009i}. We illustrate the cases $n=7,8$ here in Figure \ref{fig:A(7,8)star}. The numbering of the vertices of $\mathcal{A}^{(2m+1)\ast}$ that we use here is the same as that in \cite{evans/pugh:2010ii}, but the reverse of that used in \cite{evans/pugh:2009i}.
The relations in $A(\mathcal{A}^{(n)\ast},W)$ are
\begin{eqnarray}
& W_{112}[121] + W_{111}[111] = 0, & \nonumber \label{eqn:rel-Astar-i} \\
& W_{p-1,p,p}[p(p-1)j] + W_{p,p,p}[ppp] + W_{p,p,p+1}[p(p+1)p] = 0, & \label{eqn:rel-Astar-ii} \\
& W_{p,p,p+1}[pp(p+1)] + W_{p,p+1,p+1}[p(p+1)(p+1)] = 0, & \label{eqn:rel-Astar-iii} \\
& W_{p,p,p+1}[(p+1)pp] + W_{p,p+1,p+1}[(p+1)(p+1)p] = 0, & \label{eqn:rel-Astar-iv}
\end{eqnarray}
where $p=2,\ldots,r-1$ in (\ref{eqn:rel-Astar-ii}), and $p=1,\ldots,p'$ in (\ref{eqn:rel-Astar-iii}), (\ref{eqn:rel-Astar-iv}), where $r = \lfloor (n-1)/2 \rfloor$, $p'=r-1$ for even $n$, and $p'=r-2$ for odd $n$.
For even $n$ we have the extra relation $W_{r-1,r,r}[r(r-1)r] + W_{r,r,r}[rrr] = 0$, and for odd $n$ we have the extra relation $[r(r-1)(r-1)] = [(r-1)(r-1)r] = 0$.
We first consider the even case $n=2m+2$.
Clearly all loops $[pp]$ of length 1 are in $A/[A,A]$, $p=1,\ldots,m$.
Let $d_l^p := \mathrm{dim}(p A_l p)$. From the Hilbert series for $A$, we see that $d_{2k}^p = d_{2k+1}^p = p = d_{2m-2k-1}^p = d_{2m-2k-2}^p$ if $p \leq k$ or $p \geq m-k+1$, and $d_{2k}^p = d_{2k+1}^p = k+1 = d_{2m-2k-1}^p = d_{2m-2k-2}^p$ if $k+1 \leq p \leq m-k$, for $2k \leq m-1$. Then $\mathrm{dim}(A^S_{2k}) = \mathrm{dim}(A^S_{2k+1}) = (k+1)(m-k) = \mathrm{dim}(A^S_{2m-2k-1}) = \mathrm{dim}(A^S_{2m-2k-2})$ for $2k \leq m-1$.
Each commutator of the form $[[l(l+1)],[(l+1)l]] = [l(l+1)l] - [(l+1)l(l+1)]$ yields a relation between linearly independent paths of length 2 in $A/[A,A]$, $l=1,\ldots,m-1$. There are $m-1$ such relations, thus the dimension of $(A/[A,A])_2$ is $2(m-1) - (m-1) = m-1$. Similarly the dimension of $(A/[A,A])_3$ is $m-1$.
Each commutator of the form $C_p = [[(p-1)ppp],[p(p-1)]]$, $p=2,\ldots,m$, and $C_p' = [[(p-1)p(p+1)],[(p+1)p(p-1)]]$, $p=2,\ldots,m-1$, yield relations between linearly independent paths of length 4 in $A/[A,A]$. There is one basis path in $1 A^S_4 1$, which we may take to be $[11111]$. Let $w_1$ denote the basis element given by its image in $A/[A,A]$. Since $d_4^2 = 2$, the dimension of $2 A^S_4 2$ is 2. However, the basis can be chosen such that one of the basis paths is identified with $w_1$ in $A/[A,A]$ by $C_2$, thus we obtain one new basis path $w_2 \in (A/[A,A])_4$, which may be chosen to be $[22222]$. Similarly, the dimension of $p A^S_4 p$ is 3, $p=3,\ldots,m-2$, and the basis can be chosen such that two of the basis paths are identified with linear combinations of $w_1, w_2, \ldots, w_{p-1}$ in $A/[A,A]$ by $C_p$ and $C_p'$. Thus we obtain one new basis path $w_p \in (A/[A,A])_4$ for each $p=3,\ldots,m-2$, which may be chosen to be $[ppppp]$. The dimension of $(m-1) A^S_4 (m-1)$ is 2, but by $C_{m-1}$, $C_{m-1}'$ any such path can be identified with a linear combination of $w_1, w_2, \ldots, w_{m-2}$ in $A/[A,A]$. Similarly the single basis path in $m A^S_4 m$ can be identified with a linear combination of $w_1, w_2, \ldots, w_{m-2}$ in $A/[A,A]$. Thus we obtain a basis $\{ w_1, \ldots, w_{m-2} \}$ for $(A/[A,A])_4$.
By a similar argument, we see that the dimension of $(A/[A,A])_k$ is $m- \lfloor k/2 \rfloor$ for all $k=0,1,\ldots,2m-1$, with basis paths $[ppp \cdots p]$, for $p=1,\ldots,m- \lfloor k/2 \rfloor$.
Thus
\begin{equation}
HH_0(A) \cong S \oplus C,
\end{equation}
where the graded vector space $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{2m-1} (m- \lfloor j/2 \rfloor) t^{j}$.
We now consider the odd case $n=2m+1$.
Again, all loops $[pp]$ of length 1 are in $A/[A,A]$, this time for $p=1,\ldots,m-1$ (note that there is no edge from vertex $m$ to $m$ on $\mathcal{A}^{(2m+1)\ast}$).
From the Hilbert series for $A$, we see that $d_{2k}$ is given by the same formula as for the even case $n=2m+2$, for $2k \leq m-1$, whilst $d_{2k+1}^p = p = d_{2m-2k-2}^p$ if $p \leq k$ or $p \geq m-k$, $d_{2k+1}^p = k+1 = d_{2m-2k-3}^p$ if $k+1 \leq i \leq m-k-1$, and $d_{2k+1}^m = 0$, for $2k \leq m-2$.
Then $\mathrm{dim}(A^S_{2k}) = (k+1)(m-k) = \mathrm{dim}(A^S_{2m-2k-2})$ for $2k \leq m-1$, and $\mathrm{dim}(A^S_{2k-1}) = (k+1)(m-k-1) = \mathrm{dim}(A^S_{2m-2k-3})$ for $2k \leq m-2$.
By a similar argument as for the even case above, we see that the dimension of $(A/[A,A])_k$ is $m- \lfloor (k+1)/2 \rfloor$ for all $k=0,1,\ldots,2m-2$, with basis paths $[ppp \cdots p]$, for $p=1,\ldots,m- \lfloor (k+1)/2 \rfloor$.
Thus
\begin{equation}
HH_0(A) \cong S \oplus C,
\end{equation}
where the graded vector space $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{2m-2} (m- \lfloor (j+1)/2 \rfloor) t^{j}$.
\subsubsection{The conjugate orbifold $\mathcal{D}^{\ast}$ graphs}
The graphs $\mathcal{D}^{(n)\ast}$ are (three-colourable) unfolded versions, or $\mathbb{Z}_3$-orbifolds, of the graphs $\mathcal{A}^{(n)\ast}$, where we replace every vertex $v$ of $\mathcal{A}^{(n)\ast}$ by three vertices $v_0$, $v_1$, $v_2$, where $v_a$ is of colour $a$, such that there are edges $v_0 \rightarrow w_1$, $v_1 \rightarrow w_2$ and $v_2 \rightarrow w_0$ if and only if there is an edge $v \rightarrow w$ on $\mathcal{A}^{(n)\ast}$. The graphs $\mathcal{D}^{(7)\ast}$, $\mathcal{D}^{(8)\ast}$ are illustrated in Figure \ref{fig:D(7,8)star}.
Due to the three-colourability of the graph $\mathcal{D}^{(n)\ast}$, a closed loop on $\mathcal{A}^{(n)\ast}$ will only be a closed loop on $\mathcal{D}^{(n)\ast}$ if it has length $3k$, $k \geq 0$, and for each such closed loop on $\mathcal{A}^{(n)\ast}$, there are three corresponding closed loops of length $3k$ on $\mathcal{D}^{(n)\ast}$. However, these three closed loops are identified in $A/[A,A]$, which can be seen as follows. As in the case of $\mathcal{A}^{(n)\ast}$, $(A/[A,A])_{3k}$ is generated by paths of the form $[p_l p_{l+1} p_{l+2} p_l \cdots p_{l+3k}]$, for $l=0,1,2 \textrm{ mod } 3$ and $p=1,\ldots,r$, where $r = m- \lfloor 3k/2 \rfloor$ for $n=2m+2$ and $r = m- \lfloor (3k+1)/2 \rfloor$ for $n=2m+1$. Since $[p_l p_{l+1} p_{l+2} p_l \cdots p_{l}]$ is a cyclic permutation of $[p_{l+1} p_{l+2} p_l p_{l+1} \cdots p_{l+1}]$, we see that for $l=0,1,2 \textrm{ mod } 3$, the cyclic paths $[p_l p_{l+1} p_{l+2} p_l \cdots p_{l+3k}]$ are identified in $A/[A,A]$. Thus $(A/[A,A])_{3k}$ has a basis given by $[p_0 p_1 p_2 p_0 \cdots p_0]$, for $p=1,\ldots,r$, and $(A/[A,A])_{k'} = 0$ for $k' \not \equiv 0 \textrm{ mod } 3$.
Then
\begin{equation}
HH_0(A) \cong S \oplus C,
\end{equation}
where for $n=2m+2$ the graded vector space $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-1)/3 \rfloor} (m- \lfloor 3j/2 \rfloor) t^{3j}$,
whilst for $n=2m+1$, $H_C(t) = \sum_{j=1}^{\lfloor (2m-2)/3 \rfloor} (m- \lfloor (3j+1)/2 \rfloor) t^{3j}$.
\subsubsection{The graph $\mathcal{E}^{(8)}$ for the conformal embedding $SU(3)_5 \subset SU(6)_1$}
\begin{figure}[tb]
\begin{minipage}[t]{7.5cm}
\begin{center}
\includegraphics[width=35mm]{fig-E8}\\
\caption{Graph $\mathcal{E}^{(8)}$} \label{fig:E(8)}
\end{center}
\end{minipage}
\hfill
\begin{minipage}[t]{7.5cm}
\begin{center}
\includegraphics[width=15mm]{fig-E8star}\\
\caption{Graph $\mathcal{E}^{(8)\ast}$} \label{fig:E(8)star}
\end{center}
\end{minipage}
\end{figure}
We now consider the graph $\mathcal{E}^{(8)}$, illustrated in Figure \ref{fig:E(8)}. The unique cell system $W$ (up to equivalence) was computed in \cite[Theorem 9.1]{evans/pugh:2009i}. The quotient algebra $A$ has the relations, for $l=0,1,2$,
\begin{eqnarray*}
& [1_l 2_{l+1} 3_{l+2}] = [2_{l+1} 3_{l+2} 1_l] = [3_l 2_{l+1} 4_{l+2}] = [4_{l+2} 3_l 2_{l+1}] = 0, & \\
& \sqrt{[3]} [2_l 2_{l+1} 2_{l+2}] = -[2_l 3_{l+1} 2_{l+2}], \qquad \sqrt{[3]}[3_l 3_{l+1} 3_{l+2}] = [3_l 2_{l+1} 3_{l+2}], & \\
& \frac{-\sqrt{[3]}}{[2]} [3_l 1_{l+1} 2_{l+2}] = [3_l 2_{l+1} 2_{l+2}] + [3_l 3_{l+1} 2_{l+2}], & \\
& \frac{-\sqrt{[3]}}{[2]} [2_l 4_{l+1} 3_{l+2}] = [2_l 2_{l+1} 3_{l+2}] + [2_l 3_{l+1} 3_{l+2}]. &
\end{eqnarray*}
The only cyclic paths in $A_+$ are of the form $[2_l 2_{l+1} 2_{l+2} 2_l]$, $[3_l 3_{l+1} 3_{l+2} 3_l]$, $l=1,\ldots,3$. Now $[2_l 2_{l+1} 2_{l+2} 2_l] = [2_{l+1} 2_{l+2} 2_l 2_{l+1}]$ by cyclic permutation, and $\sqrt{[3]} [2_l 2_{l+1} 2_{l+2} 2_l] = -[2_l 3_{l+1} 2_{l+2} 2_l] = [3_{l+1} 2_{l+2} 2_l 3_{l+1}] = -[3_{l+1} 2_{l+2} 3_l 3_{l+1}] - (\sqrt{[2]}/\sqrt{[4]}) [3_{l+1} 2_{l+2} 4_l 3_{l+1}] = -\sqrt{[3]} [3_{l+1} 3_{l+2} 3_l 3_{l+1}]$ in $A/[A,A]$, where the second and last equalities follows by cyclic permutation and the others follow from the relations in $A$. Thus we see that all cyclic paths in $A_+$ are identified in $A/[A,A]$ so that
\begin{equation}
HH_0(A) \cong S \oplus \mathbb{C} [2_0 2_1 2_2 2_0].
\end{equation}
\subsubsection{The graph $\mathcal{E}^{(8)\ast}$ for the orbifold of the conformal embedding $SU(3)_5 \subset SU(6)_1 \rtimes \mathbb{Z}_3$}
Consider the graph $\mathcal{E}^{(8)\ast}$, illustrated in Figure \ref{fig:E(8)star}. The unique cell system $W$ (up to equivalence) was computed in \cite[Theorem 10.1]{evans/pugh:2009i}. The quotient algebra $A$ has the relations
\begin{eqnarray*}
& [123] = [231] = [324] = [432] = 0, \qquad \qquad [222] = \frac{-1}{\sqrt{[3]}} [232], \qquad \qquad [333] = \frac{1}{\sqrt{[3]}} [323], & \\
& \frac{-\sqrt{[3]}}{[2]} [312] = [322] + [332], \qquad \qquad \frac{-\sqrt{[3]}}{[2]} [243] = [223] + [233]. &
\end{eqnarray*}
Clearly the single edges $[22]$, $[33]$ are not in $[A,A]$, since the relations in $A$ only change paths of length $>1$, and edges are invariant under cyclic permutation. We have the relation $\sqrt{[3]} [222 \cdots 2] = -[232 \cdots 2] = -[322 \cdots 23]$ for paths of length $r$ in $A/[A,A]$, $2 \leq r \leq 5$, where the first equality follows from the relation in $A$ and the second follows by cyclic permutation.
Thus in $A/[A,A]$, for $r=2$, we obtain $\sqrt{[3]} [222] = -[323] = -\sqrt{[3]} [333]$ by the relations in $A$.
For $r=3$, we have $\sqrt{[3]} [2222] = -[3223] = [3323] +([2]/\sqrt{[3]})[3123] = \sqrt{[3]}[3333]$, by the relations in $A$, since the subpath $[123] = 0$ in $A$.
For $r=4$, we have $[3] [22222] = -\sqrt{[3]}[32223] = [32323] = \sqrt{[3]}[33323] = [3] [33333]$, by the relations in $A$, but also $\sqrt{[3]} [22222] = -[32223] = [33223] +([2]/\sqrt{[3]})[31223] = -[33233] -([2]/\sqrt{[3]})[33243] +([2]/\sqrt{[3]})[23122] = -\sqrt{[3]}[33333]$, by the relations in $A$, since the subpaths $[324] = 0 = [123]$ in $A$, and we have used the cyclic permutation relation in the penultimate equality. Then in $A/[A,A]$, we see that $[22222] = 0 = [33333]$.
For $r=5$ we have $[3] [222222] = -\sqrt{[3]}[322223] = [323223] = \sqrt{[3]}[333223] = -[3][333333]$ in $A/[A,A]$, by the relations in $A$.
Thus
\begin{equation}
HH_0(A) \cong S \oplus C,
\end{equation}
where the graded vector space $C = \mathbb{C} \{ [22], [33] \} \oplus \mathbb{C} [222] \oplus \mathbb{C} [2222] \oplus \mathbb{C} [222222]$, and has Hilbert series $H_C(t) = 2t + t^2 + t^3 + t^5$.
\subsection{Determining the Hochschild homology of $A(\mathcal{G},W)$ for trivial Nakayama automorphism} \label{Sect:HH(A)-trivial_beta}
In this section we determine the Hochschild and cyclic homology for the graphs $\mathcal{D}^{(3k)}$, $k \geq 2$, $\mathcal{A}^{(n)\ast}$, $n \geq 4$, $\mathcal{D}^{(3k)\ast}$, $k \geq 2$, and $\mathcal{E}^{(8)\ast}$.
Here the almost Calabi-Yau algebra $A$ has trivial Nakayama automorphism.
In this case, the Hochschild homology of $A$ has minimal period at most 4, thus we have $HH_i(A)^{\ast}[h] \cong HH_{3-i}(A)$, $i=1,2$, and $HH_i(A)^{\ast} \cong HH_{7-i}(A)$, $i=3,4$.
From the exactness of (\ref{seq:Connes_exact_seq}) we see that $\mathrm{ker}(B: \overline{HH}_0(A) \rightarrow \overline{HH}_1(A)) = 0$, and since the Connes differential $B$ preserves degrees, we have $\overline{HH}_1(A) \cong C \oplus X$, for some graded vector space $X$ which lives in degrees 1 to $h-2$. Then $\overline{HH}_2(A) \cong C^{\ast}[h] \oplus X^{\ast}[h]$, where $C \cong \overline{HH}_0(A)$ and $X^{\ast}[h]$ lives in degrees 2 to $h-1$. Now $B:\overline{HH}_1(A) \rightarrow \overline{HH}_2(A)$ restricts to an isomorphism $X \stackrel{\cong}{\longrightarrow} X^{\ast}[h]$ since (\ref{seq:Connes_exact_seq}) is exact, and since it preserves degrees, $X$ only lives in degrees 2 to $h-2$. A similar argument shows that $\mathrm{ker}(B: \overline{HH}_0(A) \rightarrow \overline{HH}_1(A)) \cong X^{\ast}[h]$, so that $\overline{HH}_3(A) \cong C^{\ast}[h] \oplus K'$, where the graded vector space $K'$ lives in degrees 3 to $h$, and $\overline{HH}_4(A) \cong C[h] \oplus K'^{\ast}[h]$, where $K'^{\ast}[h]$ lives in degrees $h$ to $2h-3$. Since $B:\overline{HH}_3(A) \rightarrow \overline{HH}_4(A)$ restricts to an isomorphism $K' \stackrel{\cong}{\longrightarrow} K'^{\ast}[h]$, we see that $K'$ lives only in degree $h$. We will write $K'=K[h]$ where $K$ is a vector space which lives in degree 0, so that $\overline{HH}_3(A) \cong C^{\ast}[h] \oplus K[h]$ and $\overline{HH}_4(A) \cong C[h] \oplus K^{\ast}[h]$.
Thus for any almost Calabi-Yau algebra $A$ with trivial Nakayama automorphism
$$\begin{array}{cccccccc}
&& 0 &&&& \qquad & \\
(\textrm{min deg},\textrm{max deg}) && \downarrow &&&&& \\
(0,h-3) && \overline{HH}_0(A) & \cong & C &&& \overline{HC}_0(A) \cong C \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(1,h-2) && \overline{HH}_1(A) & \cong & C & \hspace{-3mm} \oplus X && \overline{HC}_1(A) \cong X \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(2,h-1) && \overline{HH}_2(A) & \cong & C^{\ast}[h] & \hspace{-3mm} \oplus X^{\ast}[h] && \overline{HC}_2(A) \cong C^{\ast}[h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(3,h) && \overline{HH}_3(A) & \cong & C^{\ast}[h] & \hspace{-3mm} \oplus K[h] && \overline{HC}_3(A) \cong K[h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(h,2h-3) && \overline{HH}_4(A) & \cong & C[h] & \hspace{-3mm} \oplus K^{\ast}[h] && \overline{HC}_4(A) \cong C[h] \\
&& \downarrow &&&&& \\
&& \vdots &&&&& \\
\end{array}$$
where $X$ lives in degrees 2 to $h-2$, $K$ lives in degree 0, and $\overline{HH}_{5+i}(A) \cong \overline{HH}_{1+i}(A)[h]$, $\overline{HC}_{4+i}(A) \cong \overline{HC}_i(A)[h]$ for $i \geq 0$. Since $C$ is known, $X$ and $K$ can be determined from the Euler characteristic $\chi_{\overline{HC}(A)}(t)$ as they live in different degrees.
\subsubsection{The graphs $\mathcal{D}^{(3k)}$}
We consider the cases $\mathcal{D}^{(6k)}$, $\mathcal{D}^{(6k+3)}$ separately, $k \geq 1$.
For the graph $\mathcal{D}^{(6k)}$, $k \geq 1$, $\mathrm{det}(H_A(t)) = (1-t^{6k})^{4(3k(k-1)+2)/2}(1-t^{3k})/(1-t^3)^3$, thus $\chi_{\overline{HC}(A)}(t) = (\sum_{j=1}^{2k-1} 3t^{3j} - t^{3k} - (6k(k-1)+2)t^{6k})/(1-t^{6k})$.
Then since $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{2k-2} 3t^{3j} + t^{6k-3}$, we see that $H_X(t) = t^3 + \sum_{j=2}^{2k-2} 3t^{3j} + t^{3k} + t^{6k-3}$ and $H_K(t) = 6k(k-1)+2$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{D}^{(6k)},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems constructed in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C \oplus X, && HC_1(A) \cong X, \\
HH_2(A) \cong C^{\ast}[6k] \oplus X^{\ast}[6k], && HC_2(A) \cong C^{\ast}[6k], \\
HH_3(A) \cong C^{\ast}[6k] \oplus K[6k], && HC_3(A) \cong K[6k], \\
HH_4(A) \cong C[6k] \oplus K^{\ast}[6k], && HC_4(A) \cong C[6k], \\
HH_{4+i}(A) \cong HH_i(A)[6k], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[6k], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $X$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{2k-2} 3t^{3j} + t^{6k-3}$, $H_X(t) = t^3 + \sum_{j=2}^{2k-2} 3t^{3j} + t^{3k} + t^{6k-3}$ and $H_K(t)=6k(k-1)+2$ respectively, where for $k=1$, $H_X(t) = 0$.
\end{Thm}
For $\mathcal{D}^{(6k+3)}$, $k \geq 1$, $\mathrm{det}(H_A(t)) = (1-t^{6k+3})^{6k^2+3}/(1-t^3)^3$, thus $\chi_{\overline{HC}(A)}(t) = (\sum_{j=1}^{2k} 3t^{3j} - 6k^2 t^{6k+3})/(1-t^{6k+3})$. Since $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{2k-1} 3t^{3j} + t^{6k}$, we see that $H_X(t) = t^3 + \sum_{j=2}^{2k-1} 3t^{3j} + t^{6k}$ and $H_K(t) = 6k^2$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{D}^{(6k+3)},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems constructed in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C \oplus X, && HC_1(A) \cong X, \\
HH_2(A) \cong C^{\ast}[6k+3] \oplus X^{\ast}[6k+3], && HC_2(A) \cong C^{\ast}[6k+3], \\
HH_3(A) \cong C^{\ast}[6k+3] \oplus K[6k+3], && HC_3(A) \cong K[6k+3], \\
HH_4(A) \cong C[6k+3] \oplus K^{\ast}[6k+3], && HC_4(A) \cong C[6k+3], \\
HH_{4+i}(A) \cong HH_i(A)[6k+3], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[6k+3], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $X$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{2k-1} 3t^{3j} + t^{6k}$, $H_X(t) = t^3 + \sum_{j=2}^{2k-1} 3t^{3j} + t^{6k}$ and $H_K(t)=6k^2$ respectively.
\end{Thm}
\subsubsection{The $\mathcal{A}^{\ast}$ graphs}
Let $D_m$, $D_m'$ denote the determinant of the denominator $1 - \Delta_{\mathcal{G}} t + \Delta_{\mathcal{G}}^T t^2 - t^3$ of $H_A(t)$ for $\mathcal{G} = \mathcal{A}^{(2m+2)\ast}, \mathcal{A}^{(2m+1)\ast}$ respectively, where $\Delta_{\mathcal{G}}$ denotes the adjacency matrix of $\mathcal{G}$, and let $T_1 = 1-t+t^2-t^3$, $T_2 = t^2-t$.
From the properties of determinants we can deduce the recursion relations $D_m = T_1 D_{m-1} - T_2^2 D_{m-2}$ and $D_m' = (1-t^3) D_{m-1} - T_2^2 D_{m-2}$, for $m \geq 3$, and $D_1 = T_1$, $D_2 = T_1^2 - T_2^2$. It is easy to show by induction on $m$ that $D_m = (1-t)^m(1-t^{2m+2})/(1-t^2)$, and thus $D_m' = (1-t)^{m-1}(1-t^{2m+1})$.
Then for $\mathcal{A}^{(2m+2)\ast}$, $\mathrm{det}H_A(t) = (1-t^{2m+2})^m D_m^{-1} = (1-t^2)(1-t^{2m+2})^{m-1}/(1-t)^m$, thus $\chi_{\overline{HC}(A)}(t) = (mt+(m-1)t^2+mt^3+(m-1)t^4+\cdots+(m-1)t^{2m}+mt^{2m+1})/(1-t^{2m+2})$. Then $H_X(t) = 0 = H_K(t)$, i.e. $X=0=K$.
For $\mathcal{A}^{(2m+1)\ast}$, $\mathrm{det}H_A(t) = (1-t^{2m+1})^m (D_m')^{-1} = (1-t^{2m+1})^{m-1}/(1-t)^{m-1}$, thus $\chi_{\overline{HC}(A)}(t) = ((m-1)t+(m-1)t^2+(m-1)t^3+\cdots+(m-1)t^{2m})/(1-t^{2m+1})$. Then we again deduce that $H_X(t) = 0 = H_K(t)$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{A}^{(n)\ast},W)$, $n \geq 4$, where $W$ is any cell system on $\mathcal{A}^{(n)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C, && HC_1(A) = 0, \\
HH_2(A) \cong C^{\ast}[n], && HC_2(A) \cong C^{\ast}[n], \\
HH_3(A) \cong C^{\ast}[n], && HC_3(A) = 0, \\
HH_4(A) \cong C[h], && HC_4(A) \cong C[n], \\
HH_{4+i}(A) \cong HH_i(A)[n], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[n], \quad i \geq 1,
\end{array}$$
where the graded vector space $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{n-3} \lfloor (n-j-1)/2 \rfloor t^{j}$.
\end{Thm}
\subsubsection{The graph $\mathcal{D}^{(3k)\ast}$}
The Nakayama automorphism is trivial for the graphs $\mathcal{D}^{(3k)\ast}$. We consider the cases $\mathcal{D}^{(6k)\ast}$, $\mathcal{D}^{(6k+3)\ast}$ separately.
For the graph $\mathcal{D}^{(6k)\ast}$, $k \geq 1$, $\mathrm{det}(H_A(t)) = (1-t^6)(1-t^{6k})^{9k-6}/(1-t^3)^{3k-1}$, thus $\chi_{\overline{HC}(A)}(t) = ((3k-1)t^3+(3k-2)t^6+(3k-1)t^9+(3k-2)t^{12}+\cdots+(3k-1)t^{6k-3}-(6k-4)t^{6k})/(1-t^{6k})$. Then since $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-1)/3 \rfloor} (m- \lfloor 3j/2 \rfloor) t^{3j}$, we have $H_X(t) = 0$, $H_K(t) = 6k-4$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{D}^{(6k)\ast},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems constructed in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C, && HC_1(A) = 0, \\
HH_2(A) \cong C^{\ast}[6k], && HC_2(A) \cong C^{\ast}[6k], \\
HH_3(A) \cong C^{\ast}[6k] \oplus K[6k], && HC_3(A) \cong K[6k], \\
HH_4(A) \cong C[6k] \oplus K^{\ast}[6k], && HC_4(A) \cong C[6k], \\
HH_{4+i}(A) \cong HH_i(A)[6k], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[6k], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-1)/3 \rfloor} (m- \lfloor 3j/2 \rfloor) t^{3j}$ and $H_K(t)=6k-4$ respectively.
\end{Thm}
For $\mathcal{D}^{(6k+3)\ast}$, $k \geq 1$, $\mathrm{det}(H_A(t)) = (1-t^{6k+3})^{9k}/(1-t^3)^{3k}$, thus $\chi_{\overline{HC}(A)}(t) = (3kt^3+3kt^6+\cdots+3kt^{6k}-6kt^{6k+3})/(1-t^{6k+3})$. Since $C$ has Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-2)/3 \rfloor} (m- \lfloor (3j+1)/2 \rfloor) t^{3j}$, we have $H_X(t) = 0$, $H_K(t) = 6k$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{D}^{(6k+3)\ast},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems constructed in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C, && HC_1(A) \cong X, \\
HH_2(A) \cong C^{\ast}[6k+3], && HC_2(A) \cong C^{\ast}[6k+3], \\
HH_3(A) \cong C^{\ast}[6k+3] \oplus K[6k+3], && HC_3(A) \cong K[6k+3], \\
HH_4(A) \cong C[6k+3] \oplus K^{\ast}[6k+3], && HC_4(A) \cong C[6k+3], \\
HH_{4+i}(A) \cong HH_i(A)[6k+3], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[6k+3], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-2)/3 \rfloor} (m- \lfloor (3j+1)/2 \rfloor) t^{3j}$ and $H_K(t)=6k$ respectively.
\end{Thm}
\subsubsection{The graph $\mathcal{E}^{(8)\ast}$}
For the graph $\mathcal{E}^{(8)\ast}$, $\mathrm{det}(H_A(t)) = (1-t^2)(1-t^4)(1-t^8)^2/(1-t)^2$, thus $\chi_{\overline{HC}(A)}(t) = (2t+t^2+2t^3+2t^5+t^6+2t^7-2t^8)/(1-t^8)$. Then $H_X(t) = 0$, $H_K(t) = 2$, and we obtain:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{E}^{(8)\ast},W)$, where $W$ is any cell system on $\mathcal{E}^{(8)\ast}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C, && HC_1(A) = 0, \\
HH_2(A) \cong C^{\ast}[8], && HC_2(A) \cong C^{\ast}[8], \\
HH_3(A) \cong C^{\ast}[8] \oplus K[8], && HC_3(A) \cong K[8], \\
HH_4(A) \cong C[8] \oplus K^{\ast}[8], && HC_4(A) \cong C[8], \\
HH_{4+i}(A) \cong HH_i(A)[8], \quad i \geq 1, && HC_{4+i}(A) \cong HC_i(A)[8], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $K$ have Hilbert series $H_C(t) = 2t + t^2 + t^3 + t^5$ and $H_K(t)=2$ respectively.
\end{Thm}
\subsection{Determining the Hochschild homology of $A(\mathcal{G},W)$ for non-trivial Nakayama automorphism} \label{Sect:HH(A)-non-trivial_beta}
We now determine the Hochschild and cyclic homology for the graphs $\mathcal{A}^{(n)}$, $n=4,5,6,7$, $\mathcal{E}^{(8)}$.
Here the almost Calabi-Yau algebra $A$ has non-trivial Nakayama automorphism.
By a similar argument to that used in Section \ref{Sect:HH(A)-trivial_beta}, for any almost Calabi-Yau algebra $A$ with non-trivial Nakayama automorphism, we have
$$\begin{array}{cccccccc}
&& 0 &&&& \qquad & \\
(\textrm{min deg},\textrm{max deg}) && \downarrow &&&&& \\
(0,h-3) && \overline{HH}_0(A) & \cong & C &&& \overline{HC}_0(A) \cong C \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(1,h-2) && \overline{HH}_1(A) & \cong & C & \hspace{-3mm} \oplus X_1 && \overline{HC}_1(A) \cong X_1 \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(2,h-1) && \overline{HH}_2(A) & \cong & X_2 & \hspace{-3mm} \oplus X_1 && \overline{HC}_2(A) \cong X_2 \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(3,h) && \overline{HH}_3(A) & \cong & X_2 & \hspace{-3mm} \oplus K_1[h] && \overline{HC}_3(A) \cong K_1[h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(h,2h-3) && \overline{HH}_4(A) & \cong & X_3 & \hspace{-3mm} \oplus K_1[h] && \overline{HC}_4(A) \cong X_3 \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(h+1,2h-2) && \overline{HH}_5(A) & \cong & X_3 & \hspace{-3mm} \oplus X_4 && \overline{HC}_5(A) \cong X_4 \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(h+2,2h-1) && \overline{HH}_6(A) & \cong & X_3^{\ast}[3h] & \hspace{-3mm} \oplus X_4^{\ast}[3h] && \overline{HC}_6(A) \cong X_3^{\ast}[3h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(h+3,2h) && \overline{HH}_7(A) & \cong & X_3^{\ast}[3h] & \hspace{-3mm} \oplus K_1^{\ast}[2h] && \overline{HC}_7(A) \cong K_1^{\ast}[2h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(2h,3h-3) && \overline{HH}_8(A) & \cong & X_2^{\ast}[3h] & \hspace{-3mm} \oplus K_1^{\ast}[2h] && \overline{HC}_8(A) \cong X_2^{\ast}[3h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(2h+1,3h-2) && \overline{HH}_{9}(A) & \cong & X_2^{\ast}[3h] & \hspace{-3mm} \oplus X_1^{\ast}[3h] && \overline{HC}_9(A) \cong X_1^{\ast}[3h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(2h+2,3h-1) && \overline{HH}_{10}(A) & \cong & C^{\ast}[3h] & \hspace{-3mm} \oplus X_1^{\ast}[3h] && \overline{HC}_{10}(A) \cong C^{\ast}[3h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow && \hspace{-2mm} {\scriptstyle \cong} \downarrow &&& \\
(2h+3,3h) && \overline{HH}_{11}(A) & \cong & C^{\ast}[3h] & \hspace{-3mm} \oplus K_2[3h] && \overline{HC}_{11}(A) \cong K_2[3h] \\
&& \hspace{-3mm} {\scriptstyle B} \downarrow &&& \hspace{-3mm} {\scriptstyle \cong} \downarrow && \\
(3h,4h-3) && \overline{HH}_{12}(A) & \cong & C[3h] & \hspace{-3mm} \oplus K_2^{\ast}[3h] && \overline{HC}_{12}(A) \cong C[3h] \\
&& \downarrow &&&&& \\
&& \vdots &&&&& \\
\end{array}$$
where $X_1$ lives in degrees 2 to $h-2$, $X_2$ lives in degrees 3 to $h-1$, $X_3$ lives in degrees $h+1$ to $2h-3$, $X_4$ lives in degrees $h+2$ to $2h-2$, $K_i$ lives in degree 0, $i=1,2$, and $\overline{HH}_{13+i}(A) \cong \overline{HH}_{1+i}(A)[3h]$, $\overline{HC}_{12+i}(A) \cong \overline{HC}_i(A)[3h]$ for $i \geq 0$.
The graded vector space $K_1$ can be determined from the Euler characteristic $\chi_{\overline{HC}(A)}(t)$ as it is the only vector space which lives in degree $h$. The vector spaces $X_1$, $X_3$ can be determined by computing $\overline{HH}_1(A)$, $\overline{HH}_4(A)$ respectively. Then $X_2$, $X_4$, $K_2$ can each be determined from knowledge of $C \cong \overline{HH}_0(A)$, $X_1$, $X_3$ and the Euler characteristic.
\subsubsection{The $\mathcal{A}$ graphs}
Here we determine the Hochschild and cyclic homology for the graphs $\mathcal{A}^{(n)}$, $n=4,5,6,7$. The graphs $\mathcal{A}^{(n)}$, $n=5,6,7$, are illustrated in Figure \ref{fig:A(5,6,7)}. We have not yet been able to determine the Hochschild and cyclic homology for the case of general $n$.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=120mm]{fig-A5,6,7}\\
\caption{Graphs $\mathcal{A}^{(n)}$, $n=5,6,7$} \label{fig:A(5,6,7)}
\end{center}
\end{figure}
We first consider the graph $\mathcal{A}^{(4)}$, for which $\mathrm{det}(H_A(t)) = (1-t^{6})/(1-t^3)$. Thus $\chi_{\overline{HC}(A)}(t) = (t^3+t^6)/(1-t^{12})$ and we see that $H_{K_1}(t) = 0$, and since $C=0$ for all the $\mathcal{A}$ graphs, $H_{K_2}(t) = 0$. Since $\mathrm{ker}(\mu_1') \subset (V \otimes_S A)^S = 0$ and $\mathrm{ker}(\mu_4') = \mathcal{N}^S = 0$, we see that $\overline{HH}_1(A) = 0 = \overline{HH}_4(A)$. Thus $X_1 = X_3 = K_1 = 0$, and from $\chi_{\overline{HC}(A)}(t)$ we deduce that $X_2$ has Hilbert series $H_{X_2}(t) = t^3$ and $X_4 = K_2 = 0$.
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{A}^{(4)},W)$, where $W$ is any cell system on $\mathcal{A}^{(4)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S, & \qquad & HC_0(A) \cong S, \\
HH_1(A) = 0, && HC_1(A) = 0, \\
HH_2(A) \cong X, && HC_2(A) \cong X, \\
HH_3(A) \cong X, && HC_3(A) = 0, \\
HH_4(A) = 0, && HC_4(A) = 0, \\
HH_5(A) = 0, && HC_5(A) = 0, \\
HH_6(A) = 0, && HC_6(A) = 0, \\
HH_7(A) = 0, && HC_7(A) = 0, \\
HH_8(A) \cong X^{\ast}[12], && HC_8(A) \cong X^{\ast}[12], \\
HH_9(A) \cong X^{\ast}[12], && HC_9(A) = 0, \\
HH_{10}(A) = 0, && HC_{10}(A) = 0, \\
HH_{11}(A) = 0, && HC_{11}(A) = 0, \\
HH_{12}(A) = 0, && HC_{12}(A) = 0, \\
HH_{12+i}(A) \cong HH_i(A)[12], \quad i \geq 1, && HC_{12+i}(A) \cong HC_i(A)[12], \quad i \geq 1,
\end{array}$$
where the graded vector space $X$ has Hilbert series $H_{X}(t) = t^3$.
\end{Thm}
We now consider the graph $\mathcal{A}^{(5)}$, for which $\mathrm{det}(H_A(t)) = (1-t^{15})/(1-t^3)$. Thus $\chi_{\overline{HC}(A)}(t) = (t^3+t^6+t^9+t^{12})/(1-t^{15})$ and we see that $H_{K_1}(t) = 0$, and since $C=0$ for all the $\mathcal{A}$ graphs, $H_{K_2}(t) = 0$. We now explicitly determine $\overline{HH}_1(A)$ and $\overline{HH}_4(A)$.
We begin with the graded vector space $Y = \overline{HH}_1(A) = \mathrm{ker}(\mu_1')/\mathrm{Im}(\mu_2')$, and consider each graded piece $Y_j$ separately. Due to the three-colourability of $\mathcal{A}^{(5)}$, $Y_j = 0$ for $j=1,2$. Thus we only need to determine $Y_3$. A basis for $(\widetilde{V} \otimes_S A)^S_2$ is given by the elements $[2_{l+1} 1_l] \otimes [1_l 2_{l+1}]$, $[1_{l+1} 2_l] \otimes [2_l 1_{l+1}]$ and $[2_{l+1} 2_l] \otimes [2_l 2_{l+1}]$, for $l=0,1,2 \textrm{ mod } 3$.
We have $\mu_2'([2_{l+1} 1_l] \otimes [1_l 2_{l+1}]) = W_{1_l 2_{l+1} 2_{l+2}} ([2_{l+1} 2_{l+2}] \otimes [2_{l+2} 1_l 2_{l+1}] + [1_l 2_{l+1}] \otimes [2_{l+1} 2_{l+2} 1_l]) = W_{2_l 2_{l+1} 2_{l+2}} [2_{l+1} 2_{l+2}] \otimes [2_{l+2} 2_l 2_{l+1}]$, using the relations in $A$.
Thus we see that $[2_{l+1} 2_{l+2}] \otimes [2_{l+2} 2_l 2_{l+1}] = 0$ in $Y_3$, $l=0,1,2 \textrm{ mod } 3$. Thus $Y_3 = 0$ and we obtain $\overline{HH}_1(A) = 0$. Since $X_1 = 0$, we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_2$ has Hilbert series $H_{X_2}(t) = t^3$.
We now consider $Y' = \overline{HH}_4(A)$, which lives in degrees 5 to 7. Now $\mathrm{ker}(\mu_4') = \mathcal{N}^S[5]$, since $\sum_j w_j^{\ast} \beta(x) w_j = 0$ for all $x \in \mathcal{N}^S_+$ and $\mathcal{N}^S_0 = 0$. As with $\overline{HH}_1(A)$, $Y'_j = 0$ for $j=5,7$, due to the three-colourability of $\mathcal{A}^{(5)}$.
We now determine $Y'_6$. A basis for $(V \otimes_S \mathcal{N})^S_1$ is given by $[2_l 2_{l+1}] \otimes [2_{l+1}]$, $l=0,1,2 \textrm{ mod } 3$, and a basis for $\mathcal{N}^S$ is given by $[2_l 2_{l+1}]$, $l=0,1,2 \textrm{ mod } 3$. Now $\mu_5'([2_l 2_{l+1}] \otimes [2_{l+1}]) = [2_{l+1} 2_{l+2}] - [2_l 2_{l+1}]$, thus $[2_0 2_1] = [2_1 2_2] = [2_2 2_0]$ in $Y'_6$. Then $\overline{HH}_4(A) = \mathbb{C}[2_0 2_1] [5] = X_3$, and we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_4 = 0$.
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{A}^{(5)},W)$, where $W$ is any cell system on $\mathcal{A}^{(5)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S, & \qquad & HC_0(A) \cong S, \\
HH_1(A) = 0, && HC_1(A) = 0, \\
HH_2(A) \cong X_2, && HC_2(A) \cong X_2, \\
HH_3(A) \cong X_2, && HC_3(A) = 0, \\
HH_4(A) \cong X_3, && HC_4(A) \cong X_3, \\
HH_5(A) \cong X_3, && HC_5(A) = 0, \\
HH_6(A) \cong X_3^{\ast}[15], && HC_6(A) \cong X_3^{\ast}[15], \\
HH_7(A) \cong X_3^{\ast}[15], && HC_7(A) = 0, \\
HH_8(A) \cong X_2^{\ast}[15], && HC_8(A) \cong X_2^{\ast}[15], \\
HH_9(A) \cong X_2^{\ast}[15], && HC_9(A) = 0, \\
HH_{10}(A) = 0, && HC_{10}(A) = 0, \\
HH_{11}(A) = 0, && HC_{11}(A) = 0, \\
HH_{12}(A) = 0, && HC_{12}(A) = 0, \\
HH_{12+i}(A) \cong HH_i(A)[15], \quad i \geq 1, && HC_{12+i}(A) \cong HC_i(A)[15], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $X_2$ and $X_3$ have Hilbert series $H_{X_2}(t) = t^3$ and $H_{X_3}(t) = t^6$.
\end{Thm}
We now consider the graph $\mathcal{A}^{(6)}$, for which $\mathrm{det}(H_A(t)) = (1-t^6)(1-t^9)(1-t^{18})/(1-t^3)$. Thus $\chi_{\overline{HC}(A)}(t) = (t^3+t^{15}-2t^{18})/(1-t^{18})$ and we see that $H_{K_1}(t) = 0$, and since $C=0$ for all the $\mathcal{A}$ graphs, $H_{K_2}(t) = 2$. We now explicitly determine $\overline{HH}_1(A)$ and $\overline{HH}_4(A)$.
We begin with the graded vector space $Y = \overline{HH}_1(A)$. Due to the three-colourability of $\mathcal{A}^{(6)}$, $Y_j = 0$ for $j=1,2,4$, so we only need to determine $Y_3$. A basis for $(\widetilde{V} \otimes_S A)^S_2$ is given by the elements $[i_{l+1} i_l] \otimes [i_l i_{l+1}]$, $[(i+1)_2 i_1] \otimes [i_1 (i+1)_2]$, $[i_1 4_0] \otimes [4_0 i_1]$, and $[4_0 i_2] \otimes [i_2 4_0]$, for $l=0,1,2$, $i=1,2,3 \textrm{ mod } 3$.
A basis for $(V \otimes A)^S_3$ is given by $[4_0 i_1] \otimes [i_1 i_2 4_0]$, $[i_1 i_2] \otimes [i_2 4_0 i_1]$ and $[i_2 4_0] \otimes [4_0 i_1 i_2]$, for $i=1,2,3 \textrm{ mod } 3$.
Under $\mu_2'$ the basis elements of $(\widetilde{V} \otimes_S A)^S_2$ yield the following expressions after using the relations in $A$, for $i=1,2,3 \textrm{ mod } 3$:
\begin{eqnarray*}
&& \mu_2'([i_1 i_0] \otimes [i_0 i_1]) = -W_{4_0 i_1 i_2} [i_1 i_2] \otimes [i_2 4_0 i_1] = \mu_2'([i_0 i_2] \otimes [i_2 i_0]), \\
&& \mu_2'([i_2 i_1] \otimes [i_1 i_2]) = W_{4_0 i_1 i_2} [4_0 i_1] \otimes [i_1 i_2 4_0] + W_{4_0 i_1 i_2} [i_2 4_0] \otimes [4_0 i_1 i_2], \\
&& \mu_2'([(i+1)_2 i_1] \otimes [i_1 (i+1)_2]) = W_{4_0 i_1 (i+1)_2} [4_0 i_1] \otimes [i_1 (i+1)_2 4_0] \\
&& \qquad \qquad \qquad \qquad \qquad \qquad \qquad + W_{4_0 i_1 (i+1)_2} [(i+1)_2 4_0] \otimes [4_0 i_1 (i+1)_2], \\
&& \mu_2'([i_1 4_0] \otimes [4_0 i_1]) = W_{4_0 i_1 i_2} [i_2 4_0] \otimes [4_0 i_1 i_2] + W_{4_0 i_1 (i+1)_2} [(i+1)_2 4_0] \otimes [4_0 i_1 (i+1)_2] \\
&& \qquad \qquad \qquad \qquad \qquad + W_{4_0 i_1 i_2} [i_1 i_2] \otimes [i_2 4_0 i_1], \\
&& \mu_2'([4_0 i_2] \otimes [i_2 4_0]) = W_{4_0 i_1 i_2} [4_0 i_1] \otimes [i_1 i_2 4_0] + W_{4_0 (i-1)_1 i_2} [4_0 (i-1)_1] \otimes [(i-1)_1 i_2 4_0] \\
&& \qquad \qquad \qquad \qquad \qquad + W_{4_0 i_1 i_2} [i_1 i_2] \otimes [i_2 4_0 i_1].
\end{eqnarray*}
Then from $\mathrm{Im}(\mu_2')_3$ we obtain the following relations in $Y_3$: $[i_1 i_2] \otimes [i_2 4_0 i_1] = 0$ and $[4_0 i_1] \otimes [i_1 i_2 4_0] = -[i_2 4_0] \otimes [4_0 i_1 i_2] = (W_{4_0 3_1 3_2}/W_{4_0 i_1 i_2}) [4_0 3_1] \otimes [3_1 1_2 4_0]$.
We now consider $\mathrm{Ker}(\mu_1')_3$. Let $x = \sum_{i=1}^3 (\lambda_i^0 [i_2 4_0] \otimes [4_0 i_1 i_2] + \lambda_i^1 [4_0 i_1] \otimes [i_1 i_2 4_0] + \lambda_i^2 [i_1 i_2] \otimes [i_2 4_0 i_1])$ be a general element in $(V \otimes A)^S_3$. Since $\mu_1'(x) = \sum_{i=1}^3((\lambda_i^0 - \lambda_i^1) [4_0 i_1 i_2 4_0] + (\lambda_i^1 - \lambda_i^2) [i_1 i_2 4_0 i_1] + (\lambda_i^2 - \lambda_i^0) [i_2 4_0 i_1 i_2])$, then $x \in \mathrm{Ker}(\mu_1')$ if and only if $\lambda_i^0 = \lambda_i^1 = \lambda_i^2$ for each $i=1,2,3$. Using the relations from $\mathrm{Im}(\mu_2')_3$, a general element in $Y_3$ is thus of the form $\sum_i \lambda_i^0 ([i_2 4_0] \otimes [4_0 i_1 i_2] + [4_0 i_1] \otimes [i_1 i_2 4_0] + [i_1 i_2] \otimes [i_2 4_0 i_1]) = \sum_i \lambda_i^0 (-(W_{4_0 3_1 3_2}/W_{4_0 i_1 i_2}) [4_0 3_1] \otimes [3_1 1_2 4_0] + 0 + (W_{4_0 3_1 3_2}/W_{4_0 i_1 i_2}) [4_0 3_1] \otimes [3_1 1_2 4_0]) = 0$. Thus $Y_3 = 0$ and we obtain $\overline{HH}_1(A) = 0$. Since $X_1 = 0$, we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_2$ is a graded vector space with Hilbert series $H_{X_2}(t) = t^3$.
We now consider $Y' = \overline{HH}_4(A)$, which lives in degrees 6 to 9.
Now $\mathcal{N}^S_0 = \mathbb{C}[4_0]$, but $\mu_4'(\lambda [4_0]) = 2\lambda [4_0 1_1 1_2 4_0]$, thus $\lambda [4_0] \in \mathrm{ker}(\mu_4')$ if and only if $\lambda = 0$.
Thus $\mathrm{ker}(\mu_4') = \mathcal{N}^S_+[6]$, since $\sum_j w_j^{\ast} \beta(x) w_j = 0$ for all $x \in \mathcal{N}^S_+ = \mathbb{C} [4_0 1_1 1_2 4_0]$.
As with $\overline{HH}_1(A)$, $Y'_j = 0$ for $j=7,8$, due to the three-colourability of $\mathcal{A}^{(6)}$, and $Y'_6 = 0$ since $\mathrm{ker}(\mu_4')_6 = 0$.
We now determine $Y'_9$. A basis for $(V \otimes_S \mathcal{N})^S_1$ is given by $[4_0 i_1] \otimes [i_1 i_2 4_0]$, $[i_1 (i+1)_2] \otimes [(i+1)_2 4_0 (i-1)_1]$ and $[i_2 4_0] \otimes [4_0 (i-1)_1 (i-1)_2]$, $i=0,1,2 \textrm{ mod } 3$.
Now $\mu_5'([4_0 1_1] \otimes [1_1 1_2 4_0]) = -[4_0 1_1 1_2 4_0]$, thus $\mathrm{Im}(\mu_5') = \mathcal{N}^S_+$. Then $Y'_9 = 0$ and we obtain $\overline{HH}_4(A) = 0$.
Then since $X_3 = 0$, we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_4 = 0$.
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{A}^{(6)},W)$, where $W$ is any cell system on $\mathcal{A}^{(6)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S, & \qquad & HC_0(A) \cong S, \\
HH_1(A) = 0, && HC_1(A) = 0, \\
HH_2(A) \cong X, && HC_2(A) \cong X, \\
HH_3(A) \cong X, && HC_3(A) = 0, \\
HH_4(A) = 0, && HC_4(A) = 0, \\
HH_5(A) = 0, && HC_5(A) = 0, \\
HH_6(A) = 0, && HC_6(A) = 0, \\
HH_7(A) = 0, && HC_7(A) = 0, \\
HH_8(A) \cong X^{\ast}[18], && HC_8(A) \cong X^{\ast}[18], \\
HH_9(A) \cong X^{\ast}[18], && HC_9(A) = 0, \\
HH_{10}(A) = 0, && HC_{10}(A) = 0, \\
HH_{11}(A) \cong K[18], && HC_{11}(A) \cong K[18], \\
HH_{12}(A) \cong K^{\ast}[18], && HC_{12}(A) = 0, \\
HH_{12+i}(A) \cong HH_i(A)[18], \quad i \geq 1, && HC_{12+i}(A) \cong HC_i(A)[18], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $X$ and $K$ have Hilbert series $H_{X}(t) = t^3$ and $H_K(t) = 2$.
\end{Thm}
We now consider the graph $\mathcal{A}^{(7)}$, for which $\mathrm{det}(H_A(t)) = (1-t^{21})^3/(1-t^3)$. Thus $\chi_{\overline{HC}(A)}(t) = (t^3+t^6+\cdots+t^{18}-2t^{21})/(1-t^{21})$ and we see that $H_{K_1}(t) = 0$, and since $C=0$ for all the $\mathcal{A}$ graphs, $H_{K_2}(t) = 2$. We now explicitly determine $\overline{HH}_1(A)$ and $\overline{HH}_4(A)$.
We begin with the graded vector space $Y = \overline{HH}_1(A)$. Due to the three-colourability of $\mathcal{A}^{(7)}$, $Y_j = 0$ for $j=1,2,4,5$, so we only need to determine $Y_3$. A basis for $(\widetilde{V} \otimes_S A)^S_2$ is given by the elements $[1_l 3_{l-1}] \otimes [3_{l-1} 1_l]$, $[2_l 1_{l-1}] \otimes [1_{l-1} 2_l]$, $[2_l 5_{l-1}] \otimes [5_{l-1} 2_l]$, $[3_l 2_{l-1}] \otimes [2_{l-1} 3_l]$, $[3_l 4_{l-1}] \otimes [4_{l-1} 3_l]$, $[4_l 2_{l-1}] \otimes [2_{l-1} 4_l]$, $[4_l 5_{l-1}] \otimes [5_{l-1} 4_l]$, $[5_l 3_{l-1}] \otimes [3_{l-1} 5_l]$, $[5_l 4_{l-1}] \otimes [4_{l-1} 5_l]$ and $[5_l 5_{l-1}] \otimes [5_{l-1} 5_l]$, for $l=0,1,2 \textrm{ mod } 3$.
A basis for $(V \otimes A)^S_3$ is given by $[2_l 3_{l+1}] \otimes [3_{l+1} 5_{l+2} 2_l]$, $[3_l 5_{l+1}] \otimes [5_{l+1} 2_{l+2} 3_l]$, $[4_l 5_{l+1}] \otimes [5_{l+1} 2_{l+2} 4_l]$, $[5_l 2_{l+1}] \otimes [2_{l+1} 3_{l+2} 5_l]$, $[5_l 4_{l+1}] \otimes [4_{l+1} 3_{l+2} 5_l]$ and $[5_l 5_{l+1}] \otimes [5_{l+1} 4_{l+2} 5_l]$, for $l=0,1,2 \textrm{ mod } 3$.
Under $\mu_2'$ the basis elements of $(\widetilde{V} \otimes_S A)^S_2$ yield the following expressions after using the relations in $A$, for $l=0,1,2 \textrm{ mod } 3$:
\begin{eqnarray*}
\mu_2'([1_l 3_{l-1}] \otimes [3_{l-1} 1_l]) & = & W_{235} [2_{l+1} 3_{l-1}] \otimes [3_{l-1} 5_l 2_{l+1}] \;\; = \;\; \mu_2'([2_{l+1} 1_l] \otimes [1_l 2_{l+1}]), \\
\mu_2'([2_l 5_{l-1}] \otimes [5_{l-1} 2_l]) & = & W_{235} [3_{l+1} 5_{l-1}] \otimes [5_{l-1} 2_l 3_{l+1}] + W_{245} [4_{l+1} 5_{l-1}] \otimes [5_{l-1} 2_l 4_{l+1}] \\
&& \quad + W_{235} [2_{l+1} 3_{l-1}] \otimes [3_{l-1} 5_l 2_{l+1}], \\
\mu_2'([3_l 2_{l-1}] \otimes [2_{l-1} 3_l]) & = & W_{235} [5_{l+1} 2_{l-1}] \otimes [2_{l-1} 3_l 5_{l+1}] + W_{235} [3_l 5_{l+1}] \otimes [5_{l+1} 2_{l-1} 3_l], \\
\mu_2'([3_l 4_{l-1}] \otimes [4_{l-1} 3_l]) & = & W_{354} [5_{l+1} 4_{l-1}] \otimes [4_{l-1} 3_l 5_{l+1}] - W_{235} [3_l 5_{l+1}] \otimes [5_{l+1} 2_{l-1} 3_l], \\
\mu_2'([4_l 2_{l-1}] \otimes [2_{l-1} 4_l]) & = & -W_{235} [5_{l+1} 2_{l-1}] \otimes [2_{l-1} 3_l 5_{l+1}] + W_{245} [4_l 5_{l+1}] \otimes [5_{l+1} 2_{l-1} 4_l], \\
\mu_2'([4_l 5_{l-1}] \otimes [5_{l-1} 4_l]) & = & W_{455} [5_{l+1} 5_{l-1}] \otimes [5_{l-1} 4_l 5_{l+1}] - W_{245} [4_l 5_{l+1}] \otimes [5_{l+1} 2_{l-1} 4_l] \\
&& \quad - W_{235} [3_{l+1} 5_{l-1}] \otimes [5_{l-1} 2_l 3_{l+1}], \\
\mu_2'([5_l 3_{l-1}] \otimes [3_{l-1} 5_l]) & = & W_{354} [5_l 4_{l+1}] \otimes [4_{l+1} 3_{l-1} 5_l] + W_{235} [5_l 2_{l+1}] \otimes [2_{l+1} 3_{l-1} 5_l] \\
&& \quad + W_{235} [2_{l+1} 3_{l-1}] \otimes [3_{l-1} 5_l 2_{l+1}], \\
\mu_2'([5_l 4_{l-1}] \otimes [4_{l-1} 5_l]) & = & -W_{354} [5_{l+1} 4_{l-1}] \otimes [4_{l-1} 3_l 5_{l+1}] + W_{455} [5_l 5_{l+1}] \otimes [5_{l+1} 4_{l-1} 5_l] \\
&& \quad - W_{235} [5_l 2_{l+1}] \otimes [2_l 3_{l-1} 5_l], \\
\mu_2'([5_l 5_{l-1}] \otimes [5_{l-1} 5_l]) & = & -W_{245} [4_{l+1} 5_{l-1}] \otimes [5_{l-1} 2_l 4_{l+1}] - W_{354} [5_l 4_{l+1}] \otimes [4_{l+1} 3_{l-1} 5_l] \\
&& \quad - W_{455} [5_{l+1} 5_{l-1}] \otimes [5_{l-1} 4_l 5_{l+1}] - W_{455} [5_l 5_{l+1}] \otimes [5_{l+1} 4_{l-1} 5_l],
\end{eqnarray*}
where $W_{ijk} := W_{i_0 j_1 k_2} = W_{i_1 j_2 k_0} = W_{i_2 j_0 k_1}$ for vertices $i_l$, $j_l$, $k_l$ of $\mathcal{A}^{(7)}$, $l=0,1,2$.
Then from $\mathrm{Im}(\mu_2')_3$ we obtain the following relations in $Y_3$:
$$\begin{array}{l}
{} [2_{l} 3_{l+1}] \otimes [3_{l+1} 5_{l+2} 2_{l}] = 0, \quad l=0,1,2 \textrm{ mod } 3, \qquad \qquad [5_0 5_1] \otimes [5_1 4_2 5_0] = -x_2 - x_3, \\
{} [3_0 5_1] \otimes [5_1 2_2 3_0] = -(W_{245}/W_{235}) [4_0 5_1] \otimes [5_1 2_2 4_0] = -[5_1 2_2] \otimes [2_2 3_0 5_1] \\
\hspace{29.5mm} = (W_{354}/W_{235}) [5_1 4_2] \otimes [4_2 3_0 5_1] = (W_{354}/W_{235}) x_1 - (W_{455}/W_{235}) x_2, \\
{} [3_1 5_2] \otimes [5_2 2_0 3_1] = -(W_{245}/W_{235}) [4_1 5_2] \otimes [5_2 2_0 4_1] = -[5_2 2_0] \otimes [2_0 3_1 5_2] = (W_{354}/W_{235}) x_1, \\
{} [3_2 5_0] \otimes [5_0 2_1 3_2] = -(W_{245}/W_{235}) [4_2 5_0] \otimes [5_0 2_1 4_2] = -[5_0 2_1] \otimes [2_1 3_2 5_0] \\
\hspace{29.5mm} = (W_{354}/W_{235}) [5_0 4_1] \otimes [4_1 3_2 5_0] = (W_{354}/W_{235}) x_1 + (W_{455}/W_{235}) x_3,
\end{array}$$
where $x_1 = [5_2 4_0] \otimes [4_0 3_1 5_2]$, $x_2 = [5_1 5_2] \otimes [5_2 4_0 5_1]$ and $x_3 = [5_2 5_0] \otimes [5_0 4_1 5_2]$.
We now consider $\mathrm{Ker}(\mu_1')_3$. Let $x = \sum_{l=0}^2 (\lambda_1^l [3_l 5_{l+1}] \otimes [5_{l+1} 2_{l+2} 3_l] + \lambda_2^l [4_l 5_{l+1}] \otimes [5_{l+1} 2_{l+2} 4_l] + \lambda_3^l [5_l 2_{l+1}] \otimes [2_{l+1} 3_{l+2} 5_l] + \lambda_4^l [5_l 4_{l+1}] \otimes [4_{l+1} 3_{l+2} 5_l] + \lambda_5^l [5_l 5_{l+1}] \otimes [5_{l+1} 4_{l+2} 5_l])$ be a general element in $(V \otimes A)^S_3$.
Now $\mu_1'(x) = \sum_{l=0}^2 (\lambda_3^{l-1} [2_l 3_{l+1} 5_{l+2} 2_l] - \lambda_1^l [3_l 5_{l+1} 2_{l+2} 3_l] + (\lambda_1^{l-1} - \lambda_2^{l-1} (W_{235}/W_{245}) - \lambda_3^l + \lambda_4^l (W_{235}/W_{354}) + (\lambda_5^{l-1}-\lambda_5^l) (W_{235}/W_{455})) [5_l 2_{l+1} 3_{l+2} 5_l]) = 0$ if and only if $\lambda_1^l = \lambda_3^l = 0, \lambda_4^l = \lambda_2^{l-1} + (\lambda_5^l-\lambda_5^{l-1})(W_{354}/W_{455})$, for $l=0,1,2 \textrm{ mod } 3$.
Using the relations from $\mathrm{Im}(\mu_2')_3$, a general element in $Y_3$ is thus of the form $\sum_{l=0}^2 (\lambda_2^l [4_l 5_{l+1}] \otimes [5_{l+1} 2_{l+2} 4_l] + (\lambda_2^{l-1} + (\lambda_5^l-\lambda_5^{l-1})(W_{354}/W_{455})) [5_l 4_{l+1}] \otimes [4_{l+1} 3_{l+2} 5_l] + \lambda_5^l [5_l 5_{l+1}] \otimes [5_{l+1} 4_{l+2} 5_l]) = 0$.
Thus $Y_3 = 0$ and we obtain $\overline{HH}_1(A) = 0$. Since $X_1 = 0$, we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_2$ is a graded vector space with Hilbert series $H_{X_2}(t) = t^3 + t^6$.
We now consider $Y' = \overline{HH}_4(A)$, which lives in degrees 7 to 11. Now $\mathrm{ker}(\mu_4') = \mathcal{N}^S[7]$, since $\sum_j w_j^{\ast} \beta(x) w_j = 0$ for all $x \in \mathcal{N}^S_+$ and $\mathcal{N}^S_0 = 0$. As with $\overline{HH}_1(A)$, $Y'_j = 0$ for $j=7,8,10,11$, due to the three-colourability of $\mathcal{A}^{(7)}$.
We now determine $Y'_9$. A basis for $(V \otimes_S \mathcal{N})^S_2$ is given by $[4_l 5_{l+1}] \otimes [5_{l+1} 4_l]$, $[5_l 4_{l+1}] \otimes [4_{l+1} 5_l]$ and $[5_l 5_{l+1}] \otimes [5_{l+1} 5_l]$, and a basis for $\mathcal{N}^S$ is given by $[4_l 5_{l+1} 4_{l+2}]$, $[5_l 5_{l+1} 5_{l+2}]$, for $l=0,1,2 \textrm{ mod } 3$. Using the relations in $A$ we obtain $\mu_5'([4_l 5_{l+1}] \otimes [5_{l+1} 4_l]) = -(W_{555}/W_{455}) [5_{l+1} 5_{l+2} 5_l] - [4_l 5_{l+1} 4_{l+2}]$, $\mu_5'([5_l 4_{l+1}] \otimes [4_{l+1} 5_l]) = (W_{555}/W_{455}) [5_l 5_{l+1} 5_{l+2}] + [4_{l+1} 5_{l+2} 4_l]$ and $\mu_5'([5_l 5_{l+1}] \otimes [5_{l+1} 5_l]) = [5_{l+1} 5_{l+2} 5_l] - [5_l 5_{l+1} 5_{l+2}]$, for $l=0,1,2 \textrm{ mod } 3$. These yield the relations $[4_l 5_{l+1} 4_{l+2}] = -(W_{555}/W_{455}) [5_{l'} 5_{l'+1} 5_{l'+2}]$ in $Y'_9$, for all $l,l'=0,1,2 \textrm{ mod } 3$.
Thus we obtain $\overline{HH}_4(A) = \mathbb{C}[5_0 5_1 5_2] [7]$.
Then $X_3 = \mathbb{C}[5_0 5_1 5_2] [7]$ and we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_4 = 0$.
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{A}^{(7)},W)$, where $W$ is any cell system on $\mathcal{A}^{(7)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S, & \qquad & HC_0(A) \cong S, \\
HH_1(A) = 0, && HC_1(A) = 0, \\
HH_2(A) \cong X_2, && HC_2(A) \cong X_2, \\
HH_3(A) \cong X_2, && HC_3(A) = 0, \\
HH_4(A) \cong X_3, && HC_4(A) \cong X_3, \\
HH_5(A) \cong X_3, && HC_5(A) = 0, \\
HH_6(A) \cong X_3^{\ast}[21], && HC_6(A) \cong X_3^{\ast}[21], \\
HH_7(A) \cong X_3^{\ast}[21], && HC_7(A) = 0, \\
HH_8(A) \cong X_2^{\ast}[21], && HC_8(A) \cong X_2^{\ast}[21], \\
HH_9(A) \cong X_2^{\ast}[21], && HC_9(A) = 0, \\
HH_{10}(A) = 0, && HC_{10}(A) = 0, \\
HH_{11}(A) \cong K[21], && HC_{11}(A) \cong K[21], \\
HH_{12}(A) \cong K^{\ast}[21], && HC_{12}(A) = 0, \\
HH_{12+i}(A) \cong HH_i(A)[21], \quad i \geq 1, && HC_{12+i}(A) \cong HC_i(A)[21], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $X_2$, $X_3$ and $K$ have Hilbert series $H_{X_2}(t) = t^3 + t^6$, $H_{X_3}(t) = t^9$ and $H_K(t) = 2$.
\end{Thm}
\subsubsection{The graph $\mathcal{E}^{(8)}$}
For the graph $\mathcal{E}^{(8)}$, $\mathrm{det}(H_A(t)) = (1-t^6)(1-t^{12})(1-t^{24})^2/(1-t^3)^2 = \mathrm{det}(H_{A'}(t^3))$, where $A' = A(\mathcal{E}^{(8)\ast},W)$. Thus $\chi_{\overline{HC}(A)}(t) = (2t^3+t^6+2t^9+2t^{15}+t^{18}+2t^{21}-2t^{24})/(1-t^{24})$ and we see that $H_{K_1}(t) = 0$. Since $C \cong \mathbb{C} [2_0 2_1 2_2 2_0]$ which lives in degree $>0$, we see that $H_{K_2}(t) = 2$. We now explicitly determine $\overline{HH}_1(A)$ and $\overline{HH}_4(A)$.
We begin with the graded vector space $Y = \overline{HH}_1(A) = \mathrm{ker}(\mu_1')/\mathrm{Im}(\mu_2')$, and consider each graded piece $Y_j$ separately. Due to the three-colourability of $\mathcal{E}^{(8)}$, $Y_j = 0$ for $j=1,2,4,5$. We will first determine $Y_3$. A basis for $(\widetilde{V} \otimes_S A)^S_3$ is given by $[2_{l+1} 1_l] \otimes [1_l 2_{l+1}]$, $[2_{l+1} 2_l] \otimes [2_l 2_{l+1}]$, $[3_{l+1} 2_l] \otimes [2_l 3_{l+1}]$, $[4_{l+1} 2_l] \otimes [2_l 4_{l+1}]$, $[1_{l+1} 3_l] \otimes [3_l 1_{l+1}]$, $[2_{l+1} 3_l] \otimes [3_l 2_{l+1}]$, $[3_{l+1} 3_l] \otimes [3_l 3_{l+1}]$, and $[3_{l+1} 4_l] \otimes [4_l 3_{l+1}]$, for $l=1,2,3$. A basis for $(V \otimes_S A)^S_3$ is given by $[2_{l-1} 2_l] \otimes [2_l 2_{l+1} 2_{l-1}]$, $[3_{l-1} 3_l] \otimes [3_l 3_{l+1} 3_{l-1}]$, $[3_{l-1} 2_l] \otimes [2_l 2_{l+1} 3_{l-1}]$, $[2_{l-1} 3_l] \otimes [3_l 3_{l+1} 2_{l-1}]$, $[3_{l-1} 2_l] \otimes [2_l 3_{l+1} 3_{l-1}]$ and $[2_{l-1} 3_l] \otimes [3_l 2_{l+1} 2_{l-1}]$, $l=1,2,3$.
Under $\mu_2'$, $[2_{l+1} 1_l] \otimes [1_l 2_{l+1}]$ gives
\begin{eqnarray*}
\mu_2'([2_{l+1} 1_l] \otimes [1_l 2_{l+1}]) & = & \sqrt{[2][3]} ([2_{l+1}3_{l-1}] \otimes [3_{l-1}1_l2_{l+1}] + [1_l2_{l+1}] \otimes [2_{l+1}3_{l-1}1_l]) \\
& = & -\sqrt{[3][4]} ([2_{l+1}3_{l-1}] \otimes [3_{l-1}3_l2_{l+1}] + [2_{l+1}3_{l-1}] \otimes [3_{l-1}2_l2_{l+1}]),
\end{eqnarray*}
using the relations in $A$. We get the same result from considering $\mu_2'([1_l 3_{l-1}] \otimes [3_{l-1} 1_l])$. We also obtain (up to some scalar factor)
\begin{eqnarray*}
\mu_2'([3_{l-1} 2_{l+1}] \otimes [2_{l+1} 3_{l-1}]) & = & [3_l2_{l+1}] \otimes [2_{l+1}3_{l-1}3_l] + [3_{l-1}3_l] \otimes [3_l2_{l+1}3_{l-1}] \\
&& + [2_l2_{l+1}] \otimes [2_{l+1}3_{l-1}2_l] + [3_{l-1}2_l] \otimes [2_l2_{l+1}3_{l-1}],
\end{eqnarray*}
and the results for $\mu_2'([3_{l+1} 4_l] \otimes [4_l 3_{l+1}])$, $\mu_2'([4_l 2_{l-1}] \otimes [2_{l-1} 4_l])$ and $\mu_2'([2_{l-1} 3_{l+1}] \otimes [3_{l+1} 2_{l-1}])$ are given by the above results by interchanging $1_p \leftrightarrow 4_p$, $2_p \leftrightarrow 3_p$ for $p=l,l+1,l-1$. Finally, we also have (again up to some scalar factor)
\begin{eqnarray*}
\mu_2'([2_{l-1} 2_{l+1}] \otimes [2_{l+1} 2_{l-1}]) & = & [3_l2_{l+1}] \otimes [2_{l+1}2_{l-1}3_l] + [2_{l-1}3_l] \otimes [3_l2_{l+1}2_{l-1}] \\
&& + \sqrt{[3]}[2_l2_{l+1}] \otimes [2_{l+1}2_{l-1}2_l] + \sqrt{[3]}[2_{l-1}3_l] \otimes [3_l2_{l+1}2_{l-1}], \\
\mu_2'([3_{l-1} 3_{l+1}] \otimes [3_{l+1} 3_{l-1}]) & = & [2_l3_{l+1}] \otimes [3_{l+1}3_{l-1}2_l] + [3_{l-1}2_l] \otimes [2_l3_{l+1}3_{l-1}] \\
&& - \sqrt{[3]}[3_l3_{l+1}] \otimes [3_{l+1}3_{l-1}3_l] - \sqrt{[3]}[3_{l-1}2_l] \otimes [2_l3_{l+1}3_{l-1}].
\end{eqnarray*}
Then from $\mathrm{Im}(\mu_2')_3$ we obtain the following relations in $Y_3$: $[2_0 2_1] \otimes [2_1 2_2 2_0] = [2_{l-1} 2_l] \otimes [2_l 2_{l+1} 2_{l-1}] = -[3_{l-1} 3_l] \otimes [3_l 3_{l+1} 3_{l-1}] = [3_{l-1} 2_l] \otimes [2_l 2_{l+1} 3_{l-1}] = -[2_{l-1} 3_l] \otimes [3_l 3_{l+1} 2_{l-1}] = -[3_{l-1} 2_l] \otimes [2_l 3_{l+1} 3_{l-1}] = [2_{l-1} 3_l] \otimes [3_l 2_{l+1} 2_{l-1}]$, $l=1,2,3$, and thus $Y_3 = \mathbb{C} [2_0 2_1] \otimes [2_1 2_2 2_0] \cong C$.
We now determine $Y_6$. A basis for $(\widetilde{V} \otimes_S A)^S_6$ is given by $[2_{l+1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 2_{l+1}]$, $[3_{l+1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 3_{l+1}]$, $[2_{l+1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 2_{l+1}]$ and $[3_{l+1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 3_{l+1}]$, $l=1,2,3$. For $l=1,2,3$, we have (up to some scalar)
\begin{eqnarray*}
\mu_2'([2_{l+1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 2_{l+1}]) & = & [2_{l-1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 3_{l+1} 2_{l-1}] \\
&& + [2_{l+1} 2_{l-1}] \otimes [2_{l-1} 3_l 2_{l+1} 2_{l-1} 3_l 2_{l+1}], \\
\mu_2'([3_{l+1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 3_{l+1}]) & = & [3_{l-1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 2_{l+1} 3_{l-1}] \\
&& + [3_{l+1} 3_{l-1}] \otimes [3_{l-1} 2_l 3_{l+1} 3_{l-1} 2_l 3_{l+1}], \\
\mu_2'([2_{l+1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 2_{l+1}]) & = & [3_{l-1} 3_l] \otimes [3_l 2_{l+1} 3_{l-1} 3_l 2_{l+1} 3_{l-1}] \\
&& - [2_{l+1} 2_{l-1}] \otimes [2_{l-1} 3_l 2_{l+1} 2_{l-1} 3_l 2_{l+1}], \\
\mu_2'([3_{l+1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 3_{l+1}]) & = & [2_{l-1} 2_l] \otimes [2_l 3_{l+1} 2_{l-1} 2_l 3_{l+1} 2_{l-1}] \\
&& - [3_{l+1} 3_{l-1}] \otimes [3_{l-1} 2_l 3_{l+1} 3_{l-1} 2_l 3_{l+1}],
\end{eqnarray*}
which yield $\mathrm{Im}(\mu_2')_6 = (\widetilde{V} \otimes_S A)^S_6$. Thus $Y_6 = 0$, and we obtain $\overline{HH}_1(A) \cong C$. Then $X_1 = 0$ and from $\chi_{\overline{HC}(A)}(t)$ we deduce that $X_2$ is a graded vector space with Hilbert series $H_{X_2}(t) = t^3 + t^6$.
We now consider $Y' = \overline{HH}_4(A)$, which lives in degrees 8 to 13. Now $\mathrm{ker}(\mu_4') = \mathcal{N}^S[8]$, since $\sum_j w_j^{\ast} \beta(x) w_j = 0$ for all $x \in \mathcal{N}^S_+$ and $\mathcal{N}^S_0 = 0$. As with $\overline{HH}_1(A)$, $Y'_j = 0$ for $j=8,10,11,13$, due to the three-colourability of $\mathcal{E}^{(8)}$. We now determine $Y'_9$. A basis for $(V \otimes_S \mathcal{N})^S_1$ is given by $[2_l 2_{l+1}] \otimes [2_{l+1}]$ and $[3_l 3_{l+1}] \otimes [3_{l+1}]$, $l=1,2,3$, and a basis for $\mathcal{N}^S_1$ is given by $[2_l 2_{l+1}]$ and $[3_l 3_{l+1}]$, $l=1,2,3$. Now $\mu_5'([2_l 2_{l+1}] \otimes [2_{l+1}]) = [2_{l+1} 2_{l-1}] - [2_l 2_{l+1}]$ and $\mu_5'([3_l 3_{l+1}] \otimes [3_{l+1}]) = [3_{l+1} 3_{l-1}] - [3_l 3_{l+1}]$, $l=1,2,3$, thus $Y'_9 = (\mathbb{C}[2_0 2_1] \oplus \mathbb{C}[3_0 3_1])[8]$. We now determine $Y'_{12}$. A basis for $\mathcal{N}^S_4$ is given by $[2_l 3_{l+1} 2_{l-1} 2_l 2_{l+1}]$ and $[3_l 2_{l+1} 3_{l-1} 3_l 3_{l+1}]$, $l=1,2,3$. Since $\mu_5'([1_l 2_{l+1}] \otimes [2_{l+1} 3_{l-1} 3_l 1_{l+1}] = [2_{l+1} 3_{l-1} 2_l 2_{l+1} 2_{l-1}]$ up to some scalar, by using the relations in $A$, and similarly $\mu_5'([4_l 3_{l+1}] \otimes [3_{l+1} 2_{l-1} 2_l 4_{l+1}] = [3_{l+1} 2_{l-1} 3_l 3_{l+1} 3_{l-1}]$, $l=1,2,3$ we see that $Y'_{12} = 0$. Thus $\overline{HH}_4(A) = (\mathbb{C}[2_0 2_1] \oplus \mathbb{C}[3_0 3_1])[8]$, and we obtain $X_3 = (\mathbb{C}[2_0 2_1] \oplus \mathbb{C}[3_0 3_1])[8]$ and we deduce from $\chi_{\overline{HC}(A)}(t)$ that $X_4 = 0$.
To summarize:
\begin{Thm}
The Hochschild and cyclic homology of $A=A(\mathcal{E}^{(8)},W)$, where $W$ is any cell system on $\mathcal{E}^{(8)}$, is given by
$$\begin{array}{lcl}
HH_0(A) \cong S \oplus C, & \qquad & HC_0(A) \cong S \oplus C, \\
HH_1(A) \cong C, && HC_1(A) = 0, \\
HH_2(A) \cong X_2, && HC_2(A) \cong X_2, \\
HH_3(A) \cong X_2, && HC_3(A) = 0, \\
HH_4(A) \cong X_3, && HC_4(A) \cong X_3, \\
HH_5(A) \cong X_3, && HC_5(A) = 0, \\
HH_6(A) \cong X_3^{\ast}[24], && HC_6(A) \cong X_3^{\ast}[24], \\
HH_7(A) \cong X_3^{\ast}[24], && HC_7(A) = 0, \\
HH_8(A) \cong X_2^{\ast}[24], && HC_8(A) \cong X_2^{\ast}[24], \\
HH_9(A) \cong X_2^{\ast}[24], && HC_9(A) = 0, \\
HH_{10}(A) \cong C^{\ast}[24], && HC_{10}(A) \cong C^{\ast}[24], \\
HH_{11}(A) \cong C^{\ast}[24] \oplus K[24], && HC_{11}(A) \cong K[24], \\
HH_{12}(A) \cong C[24] \oplus K^{\ast}[24], && HC_{12}(A) \cong C[24], \\
HH_{12+i}(A) \cong HH_i(A)[24], \quad i \geq 1, && HC_{12+i}(A) \cong HC_i(A)[24], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $X_2$, $X_3$ and $K$ have Hilbert series $H_C(t) = t^3$, $H_{X_2}(t) = t^3 + t^6$, $H_{X_3}(t) = 2t^9$ and $H_K(t)=2$ respectively.
\end{Thm}
\section{The Hochschild cohomology of $A(\mathcal{G},W)$} \label{sect:Hoch_cohom}
\subsection{The Hochschild cohomology complex} \label{sect:Hoch_cohom-complex}
In this section we will construct a complex which determines the Hochschild cohomology of the almost Calabi-Yau algebra $A = A(\mathcal{G},W)$.
Each four-term piece of this complex will be identified up to a shift in degree with a four-term piece in the Hochschild homology complex (\ref{HH_hom_complex-almostCY}).
The Hochschild cohomology $HH^{\bullet}(A)$ of $A$ may be defined as the derived functor $HH^n(A) = \mathrm{Ext}^n_{A^e}(A,A)$, that is, the homology of the complex
$$0 \rightarrow \mathrm{Hom}_{A^e}(P_0,A) \rightarrow \mathrm{Hom}_{A^e}(P_1,A) \rightarrow \mathrm{Hom}_{A^e}(P_2,A) \rightarrow \cdots$$
where $\quad \cdots \rightarrow P_2 \rightarrow P_1 \rightarrow P_0 \rightarrow A \rightarrow 0$ is any projective resolution of $A$.
Following \cite{etingof/eu:2007}, we can make identifications $\mathrm{Hom}_{A^e}(A \otimes_S \mathcal{N}^{(k)},A) = (\mathcal{N}^{(-k)})^S$, $k=0,1,2$, by identifying $\phi \in \mathrm{Hom}_{A^e}(A \otimes_S \mathcal{N}^{(k)},A)$ with the image $\phi(1 \otimes 1) = x \in (\mathcal{N}^{(-k)})^S$. We write $\phi = x \circ - : A \otimes_S \mathcal{N}^{(-k)} \rightarrow A$, and have $\phi(y \otimes z) = x \circ (y \otimes z) = yx\beta^{k}(z)$, for $x \in (\mathcal{N}^{(-k)})^S$, $y \in A$, $z \in \mathcal{N}^{(k)}$.
We also make identifications $\mathrm{Hom}_{A^e}(A \otimes_S V \otimes_S \mathcal{N}^{(k)},A) = (\widetilde{V} \otimes_S \mathcal{N}^{(-k)})^S[-2]$, $k=0,1,2$, by identifying $\phi \in \mathrm{Hom}_{A^e}(A \otimes_S V \otimes_S \mathcal{N}^{(k)},A)$ which maps $1 \otimes a \otimes 1 \mapsto x_a$ with the element $\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a \in (\widetilde{V} \otimes_S \mathcal{N}^{(-k)})^S$. We write $\phi = \sum_{b \in \mathcal{G}_1} \widetilde{b} \otimes x_b \circ - : A \otimes_S V \otimes_S \mathcal{N}^{(-k)} \rightarrow A$, and have $\phi(y \otimes a \otimes z) = \sum_{b \in \mathcal{G}_1} \widetilde{b} \otimes x_b \circ (y \otimes a \otimes z) = yx_a\beta^{k}(z)$, for $\widetilde{a} \otimes x_a \in (\widetilde{V} \otimes_S \mathcal{N}^{(-k)})^S$, $y \in A$, $z \in \mathcal{N}^{(k)}$.
Similarly, we identify $\mathrm{Hom}_{A^e}(A \otimes_S \widetilde{V} \otimes_S \mathcal{N}^{(k)},A) = (V \otimes_S \mathcal{N}^{(-k)})^S[-2]$, $k=0,1,2$, by identifying $\phi$ which maps $1 \otimes \widetilde{a} \otimes 1 \mapsto y_a$ with the element $\sum_{a \in \mathcal{G}_1} a \otimes y_a$. We write $\phi = \sum_{b \in \mathcal{G}_1} b \otimes y_b \circ - : A \otimes_S \widetilde{V} \otimes_S \mathcal{N}^{(-k)} \rightarrow A$, and have $\phi(y \otimes \widetilde{a} \otimes z) = yy_a\beta^{k}(z)$, for $a \otimes y_a \in (V \otimes_S \mathcal{N}^{(-k)})^S$, $y \in A$, $z \in \mathcal{N}^{(k)}$.
Applying the functor $\mathrm{Hom}_{A^e}(-,A)$ to the periodic resolution (\ref{resolution-almostCY}) we get the Hochschild cohomology complex:
\begin{eqnarray}
&& (\mathcal{N}^{(2)})^S[-h] \stackrel{\mu_{4}^{\ast}}{\leftarrow} A^S[-3] \stackrel{\mu_{3}^{\ast}}{\leftarrow} (V \otimes_S A)^S[-3] \stackrel{\mu_{2}^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S A)^S[-2] \stackrel{\mu_{1}^{\ast}}{\leftarrow} A^S \leftarrow 0 \nonumber \\
&& \mathcal{N}^S[-2h] \stackrel{\mu_{8}^{\ast}}{\leftarrow} (\mathcal{N}^{(2)})^S[-h-3] \stackrel{\mu_{7}^{\ast}}{\leftarrow} (V \otimes_S \mathcal{N}^{(2)})^S[-h-3] \stackrel{\mu_{6}^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S \mathcal{N}^{(2)})^S[-h-2] \stackrel{\mu_{5}^{\ast}}{\leftarrow} \nonumber \\
& \cdots & \leftarrow A^S[-3h] \stackrel{\mu_{12}^{\ast}}{\leftarrow} \mathcal{N}^S[-2h-3] \stackrel{\mu_{11}^{\ast}}{\leftarrow} (V \otimes_S \mathcal{N})^S[-2h-3] \stackrel{\mu_{10}^{\ast}}{\leftarrow} (\widetilde{V} \otimes_S \mathcal{N})^S[-2h-2] \stackrel{\mu_{9}^{\ast}}{\leftarrow} \label{HH_cohom_complex-almostCY}
\end{eqnarray}
\begin{Prop}
We have $\mu_i^{\ast} = \pm\mu_{16-i}'$.
\end{Prop}
\noindent \emph{Proof}:
(i) $\mu_1^{\ast} = -\mu_3'$:
Let $a \in V$ and $x \in A^S$. Then
$$\mu_1^{\ast}(x)(1 \otimes a \otimes 1) = x \circ \mu_1(1 \otimes a \otimes 1) = x \circ (a \otimes 1 - 1 \otimes a) = ax-xa.$$
So $\mu_1^{\ast}(x)$ maps $1 \otimes a \otimes 1 \mapsto [a,x]$, giving $\mu_1^{\ast}(x) = \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes [a,x] = -\mu_3'(x)$.
Similarly, $\mu_5^{\ast}(x)$ maps $1 \otimes a \otimes 1 \mapsto ax-x\beta(a)$, giving $\mu_5^{\ast}(x) = \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes (ax-x\beta(a)) = -\mu_{11}'(x)$, and we also have $\mu_9^{\ast}(x) = \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes (ax-x\beta^2(a)) = -\mu_{7}'(x)$. \\
(ii) $\mu_2^{\ast} = \mu_2'$:
Let $a' \in V$ and for each $a \in V$ let $x_a$ be a homogeneous element in $A$ such that $\widetilde{a} \otimes x_a \in (\widetilde{V} \otimes A)^S$. Then
\begin{eqnarray*}
\lefteqn{\mu_2^{\ast}(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a)(1 \otimes \widetilde{a'} \otimes 1) \;\; = \;\; \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a \circ \mu_2(1 \otimes \widetilde{a'} \otimes 1)} \\
& = & \sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a \circ \left( \sum_{b,b' \in \mathcal{G}_1} W_{a'bb'} (b \otimes b' \otimes 1 + 1 \otimes b \otimes b') \right) \;\; = \;\; \sum_{b,b' \in \mathcal{G}_1} W_{a'bb'} (bx_{b'}+x_{b}b').
\end{eqnarray*}
So $\mu_2^{\ast}(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a)$ maps $1 \otimes \widetilde{a'} \otimes 1 \mapsto \sum_{b,b' \in \mathcal{G}_1} W_{a'bb'} (bx_{b'}+x_{b}b')$, giving $\mu_2^{\ast}(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a) = \sum_{a,b,b' \in \mathcal{G}_1} W_{abb'} (a \otimes bx_{b'} + a \otimes x_{b}b') = \mu_2'(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a)$.
Similarly, $\mu_6^{\ast}(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a) = \sum_{a,b,b' \in \mathcal{G}_1} W_{abb'} (a \otimes bx_{b'} + a \otimes x_{b}\beta(b')) = \mu_{10}'(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a)$ and $\mu_{10}^{\ast}(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a) = \sum_{a,b,b' \in \mathcal{G}_1} W_{abb'} (a \otimes bx_{b'} + a \otimes x_{b}\beta^2(b')) = \mu_{6}'(\sum_{a \in \mathcal{G}_1} \widetilde{a} \otimes x_a)$. \\
(iii) $\mu_3^{\ast} = -\mu_1'$:
For each $a \in V$ let $y_a$ be a homogeneous element in $A$ such that $a \otimes y_a \in (V \otimes A)^S$. Then
\begin{eqnarray*}
\mu_3^{\ast}(\sum_{a \in \mathcal{G}_1} a \otimes y_a)(1 \otimes 1) & = & \sum_{a \in \mathcal{G}_1} a \otimes y_a \circ \mu_3(1 \otimes 1) \\
& = & \sum_{a \in \mathcal{G}_1} a \otimes y_a \circ \sum_{b \in \mathcal{G}_1} (b \otimes \widetilde{b} \otimes 1 - 1 \otimes \widetilde{b} \otimes b) \;\; = \;\; \sum_{b \in \mathcal{G}_1} (by_{b}-y_{b}b).
\end{eqnarray*}
So $\mu_3^{\ast}(\sum_{a \in \mathcal{G}_1} a \otimes y_a)$ maps $1 \otimes 1 \mapsto \sum_{b \in \mathcal{G}_1} [b,y_b]$, giving $\mu_3^{\ast}(\sum_{a \in \mathcal{G}_1} a \otimes y_a) = \sum_{a \in \mathcal{G}_1} [a,y_a] = -\mu_1'(\sum_{a \in \mathcal{G}_1} a \otimes y_a)$.
Similarly, $\mu_7^{\ast}(\sum_{a \in \mathcal{G}_1} a \otimes y_a) = \sum_{a \in \mathcal{G}_1} (ay_{a}-y_{a}\beta(a)) = -\mu_9'(\sum_{a \in \mathcal{G}_1} a \otimes y_a)$ and $\mu_{11}^{\ast}(\sum_{a \in \mathcal{G}_1} a \otimes y_a) = \sum_{a \in \mathcal{G}_1} (ay_{a}-y_{a}\beta^2(a)) = -\mu_5'(\sum_{a \in \mathcal{G}_1} a \otimes y_a)$. \\
(iv) $\mu_4^{\ast} = \mu_{12}'$:
Let $x \in A^S$. Then
$$\mu_4^{\ast}(x)(1 \otimes 1) = x \circ \mu_4(1 \otimes 1) = x \circ \sum_j w_j \otimes w_j^{\ast} = \sum_j w_j x w_j^{\ast},$$
where $\{ w_j \}$ is a homogeneous basis for $A$ and $\{ w_j^{\ast} \}$ is its corresponding dual basis.
So $\mu_4^{\ast}(x)$ maps $1 \otimes 1 \mapsto \sum_j w_j x w_j^{\ast}$, giving $\mu_4^{\ast}(x) = \sum_j w_j x w_j^{\ast} = \mu_{12}'(x)$. Similarly, $\mu_8^{\ast}(x) = \sum_j w_j x \beta(w_j^{\ast}) = \mu_{8}'(x)$ and $\mu_{12}^{\ast}(x) = \sum_j w_j x \beta^2(w_j^{\ast}) = \mu_4'(x)$.
\hfill
$\Box$
Thus we see that we can identify, up to a shift in degree, each four-term portion of the cohomology complex (\ref{HH_cohom_complex-almostCY}) with a portion of the homology complex (\ref{HH_hom_complex-almostCY}):
\begin{eqnarray*}
HH^i(A) & \cong & HH_{3-i}(A)[-3], \qquad \qquad \quad i=1,2, \\
HH^i(A) & \cong & HH_{15-i}(A)[-3h-3], \qquad \; i=3,\ldots,12, \\
HH^{12+i}(A) & \cong & HH^i(A)[-3h], \qquad \qquad \quad \; i=1,2,\ldots,
\end{eqnarray*}
and the self-duality of the homology complex (\ref{HH_hom_complex-almostCY}) yields the relations
\begin{eqnarray*}
HH^i(A)^{\ast} & \cong & HH^{7-i}(A), \qquad \qquad i=1,\ldots,6, \\
HH^i(A)^{\ast} & \cong & HH^{19-i}(A), \qquad \qquad i=7,\ldots,11.
\end{eqnarray*}
\subsection{The Hochschild cohomology of $A = A(\mathcal{G},W)$} \label{sect:Hoch_cohom-results}
For $HH^0(A) = \mathrm{ker}(\mu_1^{\ast})/\mathrm{Im}(\mu_0^{\ast}) = \mathrm{ker}(\mu_1^{\ast})$, we have $HH^0(A) \cong HH_3(A)'[-3] \oplus L$, where $HH_3(A)' = \oplus_{j=3}^{h-1} HH_3(A)_j$ and $L = \mathbb{C} \{ u_{j\nu(j)} | \, \nu(j) = j \}$.
Then we have the following results for the Hochschild cohomology of $A$:
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{A}^{(4)},W)$, where $W$ is any cell system on $\mathcal{A}^{(4)}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong X[-3], & \qquad & HH^1(A) \cong X[-3], \\
HH^6(A) \cong X^{\ast}[-3], && HH^7(A) \cong X^{\ast}[-3], \\
HH^{12}(A) \cong X[-15], && HH^{j}(A) = 0, \quad j = 2,\ldots,5,8,\ldots,11,
\end{array}$$
and $HH^{12+i}(A) \cong HH^i(A)[-12]$ for $i \geq 1$, where the graded vector space $X$ has Hilbert series $H_X(t) = t^3$.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{A}^{(5)},W)$, where $W$ is any cell system on $\mathcal{A}^{(5)}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong X_2[-3], & \qquad & HH^1(A) \cong X_2[-3], \\
HH^2(A) = 0, && HH^3(A) = 0, \\
HH^4(A) = 0, && HH^5(A) = 0, \\
HH^6(A) \cong X_2^{\ast}[-3], && HH^7(A) \cong X_2^{\ast}[-3], \\
HH^8(A) \cong X_3^{\ast}[-3], && HH^9(A) \cong X_3^{\ast}[-3], \\
HH^{10}(A) \cong X_3[-18], && HH^{11}(A) \cong X_3[-18], \\
HH^{12}(A) \cong X_2[-18], && HH^{12+i}(A) \cong HH^i(A)[-15], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $X_2$ and $X_3$ have Hilbert series $H_{X_2}(t) = t^3$ and $H_{X_3}(t) = t^6$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{A}^{(6)},W)$, where $W$ is any cell system on $\mathcal{A}^{(6)}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong X[-3] \oplus L, & \qquad & HH^1(A) \cong X[-3], \\
HH^2(A) = 0, && HH^3(A) \cong K^{\ast}[-3], \\
HH^4(A) \cong K[-3], && HH^5(A) = 0, \\
HH^6(A) \cong X^{\ast}[-3], && HH^7(A) \cong X^{\ast}[-3], \\
HH^8(A) = 0, && HH^9(A) = 0, \\
HH^{10}(A) = 0, && HH^{11}(A) = 0, \\
HH^{12}(A) \cong X[-21], && HH^{12+i}(A) \cong HH^i(A)[-18], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $L$, $X$ and $K$ have Hilbert series $H_L(t) = t^3$, $H_{X}(t) = t^3$ and $H_K(t)=2$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{A}^{(7)},W)$, where $W$ is any cell system on $\mathcal{A}^{(7)}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong X_2[-3], & \qquad & HH^1(A) \cong X_2[-3], \\
HH^2(A) = 0, && HH^3(A) \cong K^{\ast}[-3], \\
HH^4(A) \cong K[-3], && HH^5(A) = 0, \\
HH^6(A) \cong X_2^{\ast}[-3], && HH^7(A) \cong X_2^{\ast}[-3], \\
HH^8(A) \cong X_3^{\ast}[-3], && HH^9(A) \cong X_3^{\ast}[-3], \\
HH^{10}(A) \cong X_3[-24], && HH^{11}(A) \cong X_3[-24], \\
HH^{12}(A) \cong X_2[-24], && HH^{12+i}(A) \cong HH^i(A)[-21], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $X_2$, $X_3$ and $K$ have Hilbert series $H_{X_2}(t) = t^3 + t^6$, $H_{X_3}(t) = t^9$ and $H_K(t)=2$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{D}^{(6k)},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems given in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[6k-3] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[6k-3] \oplus X^{\ast}[6k-3], \\
HH^2(A) \cong C[-3] \oplus X[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^{4+i}(A) \cong HH^i(A)[-6k], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$, $X$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{2k-2} 3t^{3j} + t^{6k-3}$, $H_L(t) = (3k(2k-1)+3)t^{6k-3}$, $H_X(t) = t^3 + \sum_{j=2}^{2k-2} 3t^{3j} + t^{3k} + t^{6k-3}$ and $H_K(t)=6k(k-1)+2$ respectively, where for $k=1$, $H_X(t) = 0$.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{D}^{(6k+3)},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems given in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[6k] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[6k] \oplus X^{\ast}[6k], \\
HH^2(A) \cong C[-3] \oplus X[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^{4+i}(A) \cong HH^i(A)[-6k-3], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$, $X$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{2k-1} 3t^{3j} + t^{6k}$, $H_L(t) = (3k(2k+1)+3)t^{6k}$, $H_X(t) = t^3 + \sum_{j=2}^{2k-1} 3t^{3j} + t^{6k}$ and $H_K(t)=6k^2$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{A}^{(n)\ast},W)$, $n \geq 4$, where $W$ is any cell system on $\mathcal{A}^{(n)\ast}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[n-3] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[n-3], \\
HH^2(A) \cong C[-3], && HH^3(A) \cong C[-3], \\
HH^4(A) \cong C^{\ast}[-3], && HH^{4+i}(A) \cong HH^i(A)[-n], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$ have Hilbert series $H_C(t) = \sum_{j=1}^{n-3} \lfloor (n-j-1)/2 \rfloor t^{j}$ and $H_L(t) = \lfloor (n-1)/2 \rfloor t^{n-3}$.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{D}^{(6k)\ast},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems given in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[6k-3] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[6k-3], \\
HH^2(A) \cong C[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^{4+i}(A) \cong HH^i(A)[-6k], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-1)/3 \rfloor} (m- \lfloor 3j/2 \rfloor) t^{3j}$, $H_L(t) = (9k-3)t^{6k-3}$ and $H_K(t)=6k-4$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{D}^{(6k+3)},W)$, $k \geq 1$, where $W$ is equivalent to one of the cell systems given in \cite{evans/pugh:2009i}, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[6k] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[6k], \\
HH^2(A) \cong C[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^{4+i}(A) \cong HH^i(A)[-6k-3], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$, $K$ have Hilbert series $H_C(t) = \sum_{j=1}^{\lfloor (2m-2)/3 \rfloor} (m- \lfloor (3j+1)/2 \rfloor) t^{3j}$, $H_L(t) = (9k+3)t^{6k}$ and $H_K(t)=6k$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{E}^{(8)},W)$, where $W$ is any cell system on $\mathcal{E}^{(8)}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong X_2[-3], & \qquad & HH^1(A) \cong X_2[-3], \\
HH^2(A) \cong C[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^5(A) \cong C^{\ast}[-3], \\
HH^6(A) \cong X_2^{\ast}[-3], && HH^7(A) \cong X_2^{\ast}[-3], \\
HH^8(A) \cong X_3^{\ast}[-3], && HH^9(A) \cong X_3^{\ast}[-3], \\
HH^{10}(A) \cong X_3[-27], && HH^{11}(A) \cong X_3[-27], \\
HH^{12}(A) \cong X_2[-27], && HH^{12+i}(A) \cong HH^i(A)[-24], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $X_2$, $X_3$ and $K$ have Hilbert series $H_C(t) = t^3$, $H_{X_2}(t) = t^3 + t^6$, $H_{X_3}(t) = 2t^9$ and $H_K(t)=2$ respectively.
\end{Thm}
\begin{Thm}
The Hochschild cohomology of $A=A(\mathcal{E}^{(8)\ast},W)$, where $W$ is any cell system on $\mathcal{E}^{(8)\ast}$, is given by
$$\begin{array}{lcl}
HH^0(A) \cong C^{\ast}[5] \oplus L, & \qquad & HH^1(A) \cong C^{\ast}[5], \\
HH^2(A) \cong C[-3], && HH^3(A) \cong C[-3] \oplus K^{\ast}[-3], \\
HH^4(A) \cong C^{\ast}[-3] \oplus K[-3], && HH^{4+i}(A) \cong HH^i(A)[-8], \quad i \geq 1,
\end{array}$$
where the graded vector spaces $C$, $L$, $K$ have Hilbert series $H_C(t) = 2t + t^2 + t^3 + t^5$, $H_L(t) = 4t^5$ and $H_K(t)=2$ respectively.
\end{Thm}
\paragraph{Acknowledgements}
Both authors were supported by the Marie Curie Research Training Network MRTN-CT-2006-031962 EU-NCG.
The authors would like to thank Karin Erdmann, Pavel Etingof, Victor Ginzburg and Jean-Louis Loday for helpful discussions and correspondence.
|
\section{\label{sec:level1}First-level heading:\protect\\ The line
\section{Introduction}
Efficient and faithful storage of quantum states of light lies at the heart of long distance quantum communication \cite{Duan_2001,Sangouard_2009b} and remarkable progress and achievements have been done in recent years \cite{Hammerer_2010,Simon_2010}. A device allowing such storage, the so-called quantum memory, can be implemented in various physical systems, which can be either atomic ensembles \cite{Hammerer_2010} or single atom like systems (typically single atoms or ions \cite{Specht_2011, Stute_2011}, quantum dots \cite{Yilmaz_2010}, superconducting qubits \cite{Devoret_2004} or NV centers \cite{Dutt_2007}). In the present article we concentrate on the latter situation and consider single atom systems, where much experimental progress has been done recently. The demanding part in these systems is that a strong coupling between the atom and the light is required. In the atomic system, this can be achieved by using high numerical aperture optical elements \cite{Tey_2008, Tey_2009, Chen_2011} or a high finesse cavity \cite{Specht_2011}, where the quantum memory application has already been demonstrated using a mapping of the polarization of the light qubit onto a single $^{87}{\rm Rb}$ atom. Single atom systems are also well suited for creating and manipulating the quantum information experiments demonstrating entanglement generation between two individual atoms \cite{Wilk_2010} and quantum gate operations between neutral atoms \cite{Isenhower_2010} and ions \cite{Benhelm_2008} have been realized. Moreover, for long lived information storage, one usually needs to transfer the optical coherence into the coherence between ground states. This is usually achieved using another strong laser beam between the excited state and the state used for storage.
In this paper, we propose a quantum memory setup consisting of a single two-level atom in a half cavity, in which we allow for an arbitrary motion of the mirror to modify the atom-light interaction --- a natural extension of the previous work done by other authors \cite{Eschner_2001, Dorner_2002, Wilson_2003, Glaetzle_2010} and \cite{Green_2011}. We show by explicit calculation that various temporal shapes of the input single photon pulse can be efficiently stored by the atom-mirror system, provided the motion of the mirror is optimized. A feature of this scheme is that there is no need for an additional atomic level nor the strong control/transfer laser. We discuss the memory efficiency and fidelity as well as possible implementations, such as a single atom/ion in a half cavity or a superconducting qubit coupled to a 1D transmission line terminated by a SQUID.
The paper is organized as follows: we present a derivation of the optimized time-dependent decay rate that maximizes the efficiency of the storage in Sec. \ref{sec QM model}. We illustrate these results with an example of an input single photon time-bin qubit and discuss possible experimental realizations of the quantum memory scheme in Sec. \ref{sec Sim_Imp}.\\
~\\
\section{The quantum memory model}
\label{sec QM model}
\subsection{General optical Bloch equations}
\label{sec obe}
We study a single two-level atom sitting in front of a moving mirror (see Figure {\ref{fig setup}}). The incident pulse propagates along the $z-$axis and first interacts with the atom. The positive frequency part of the continuum electric field operator in the standing wave basis and the interaction picture reads \cite{Blow_1990, Dorner_2002, Domokos_2002}
\begin{figure}[b]
\includegraphics[scale=0.35]{QM_setup.pdf}
\caption{ Sketch of the quantum memory setup: an arbitrary single photon wave packet interacts with a two-level atom which has a initial distance $L$ from the movable mirror, whose motion is described by $l(t)$. $\gamma_p$ and $\gamma'$ describe the decay rates into the pulse mode and the environment, respectively (remark: in the implementation that we consider, the pulse durations are much longer than $L/c$).}
\label{fig setup}
\end{figure}
\begin{equation}
\label{eq E+}
\hat{\bm E}^{(+)}(z,t)=i \sum_{\lambda} \int_0^{\infty} {\rm d} \omega \,\, A(\omega) {\bm \epsilon}_{\omega,\lambda} \sin (kz) \,e^{-i\omega t} \hat{a}_{\omega}(t),
\end{equation}
\noindent where $k=\omega/c$, $c$ is the vacuum speed of light, ${\bm \epsilon}_{\omega,\lambda}$ with $\lambda=\{1, 2\}$ denotes the unit polarization of mode $\omega$ and the coefficient $A(\omega)$ accounts for the correct normalization of the electric field (i.e. that the total energy of a single photon Fock state of frequency $\omega_0$ is $\hbar \omega_0$). We denote the initial distance between the atom and the mirror by $L$. The main goal of this article is to investigate the dynamics of a two-level atom and a pulse in front of a moving mirror. The dynamics is given by the time-dependent decay rate, which reaches its minimal value 0 for an atom sitting at the node and maximal value $2\gamma_0$ for the atom at the antinode of the cavity, a well known result from a quantum cavity electrodynamics. Here, we denoted by $\gamma_0$ the atomic decay rate in free space. We describe the motion of the mirror by a time-dependent function $l(t)$, such that the atom-mirror distance is given by $L-l(t)$ for any time $t$.
The atomic diploe operator in interaction picture reads
\begin{equation}
\hat{\bm d} = \bm d \left(\hat{\sigma}_- e^{-i\omega_a t} + \hat{\sigma}_+ e^{i\omega_a t} \right),
\label{eq d}
\end{equation}
\noindent where $\omega_a$ is the atomic transition frequency and $\hat{\sigma}_+=\ket{e}\bra{g}, \hat{\sigma}_-=\ket{g}\bra{e}, \hat{\sigma}_z=\ket{e}\bra{e}-\ket{g}\bra{g}=\hat{\sigma}_{+}\hat{\sigma}_{-}-\hat{\sigma}_{-}\hat{\sigma}_{+}$ are the usual two-level atom operators with a ground and excited states $\ket{g}$ and $\ket{e}$.
The dipole interaction Hamiltonian is equal to the scalar product of the atomic dipole (Eq. {\ref{eq d}}) and the electric field (the positive part of which is given by Eq. {\ref{eq E+}}), $\hat{H}_I = -\hat{\bm d} \cdot \hat{\bm E}$. The Hamiltonian in the interaction picture, after making the rotating wave approximation, is given by
\begin{equation}
\hat{H}_I(t)=-i \hbar \sum_{\lambda} \int {\rm d} \omega [g_{\omega, \lambda} \hat{\sigma}_{+} \hat{a}_{\omega} \sin\left[k(L-l(t))\right]\, e^{-i(\omega-\omega_a)t}-h.c.].
\end{equation}
In the following, we assume that the atomic dipole ${\bm d}$ is oriented parallel to the polarization of the field $\bm \epsilon$ and thus yielding the maximized coupling
\begin{equation}
g_{\omega} \equiv g_{\omega, \lambda} = \frac{{d}\,A(\omega)}{\hbar},
\end{equation}
\noindent where is $d=|{\bm d}|$ is the scalar atomic dipole momentum.
The Heisenberg equations of motion of the field and atomic operators are
\begin{eqnarray}
\label{eq_dd ef1}
\dot{\hat{a}}_{\omega} &=& g^*_{\omega} \sin\left[k(L-l(t))\right]\,e^{i(\omega-\omega_a)t} \,\hat{\sigma}_{-},
\end{eqnarray}
\vskip -0.5cm
\begin{eqnarray}
\label{eq_dd_am1}
\dot{\hat{\sigma}}_{-} &=& -\frac{\gamma'}{2}\hat{\sigma}_{-} +\hat{\zeta}_-\\ \nonumber
&+&\hat{\sigma}_{z}\int {\rm d} \omega g_{\omega} \hat{a}_{\omega} \sin\left[k(L-l(t))\right]\,e^{-i(\omega-\omega_a)t},
\end{eqnarray}
\vskip -0.5cm
\begin{eqnarray}
\label{eq_dd_az1}
\dot{\hat{\sigma}}_{z} &=& -\gamma'(\hat{\sigma}_{z}+1)+\hat{\zeta}_z \\ \nonumber
&-& 2 \int {\rm d} \omega \sin\left[k(L-l(t))\right] \,[ g_{\omega} \hat{\sigma}_{+} \hat{a}_{\omega}\,e^{-i(\omega-\omega_a)t}+h.c.],
\end{eqnarray}
\noindent in which the decay term $\gamma'$ and the noise operators $\hat{\zeta}$ are introduced to account for the interaction of the atom with the environment. The explicit form of the noise operator is discussed in \cite{Wang_2011}. Moreover, as a consequence of our initial conditions, the noise operators do not come into play, as explained further in the text.
By integrating Eq.(\ref{eq_dd ef1}), we can separate the field operator into two parts:
\begin{equation}
\label{eq_dd_ef2}
\hat{a}_{\omega}(t)=\hat{a}_{\omega}(t_0)+ \int_{t_0}^t {\rm d} t'\, g^*_{\omega}\sin\left[k(L-l(t'))\right]\,e^{i(\omega-\omega_a)t'}\,{\hat{\sigma}_{-}(t')},
\end{equation}
\noindent where the first term refers to the initial field having evolved freely from $t_0$ to $t$ and the second term is the field created by the atomic dipole during the time period $t-t_0$. These contributions are usually called the ``free field" and the ``source field".
After substituting Eq.(\ref{eq_dd_ef2}) into Eq.(\ref{eq_dd_am1}) and Eq.(\ref{eq_dd_az1}), we get the modified optical Bloch equations
\begin{eqnarray}
\label{eq_dd_am2}
\dot{\hat{\sigma}}_{-}(t) &=& -\frac{\gamma'}{2}\hat{\sigma}_{-}(t) +\hat{\zeta}_- \\ \nonumber
&+& \hat{\sigma}_{z}(t)\int {\rm d} \omega \, g_{\omega}\,\sin\left[k(L-l(t))\right]\,e^{-i(\omega-\omega_a)t}\, \hat{a}_{\omega}(t_0) \\ \nonumber
&+& \hat{\sigma}_{z}(t)\int_{t_0}^{t} {\rm d} t'\,\int {\rm d} \omega \,{|g_{\omega}|}^2 \hat{\sigma}_{-}(t') \\ \nonumber
&\times& \sin\left[k(L-l(t))\right] \,\sin[k(L-l(t'))]\,e^{-i(\omega-\omega_a)(t-t')},
\end{eqnarray}
\begin{eqnarray}
\label{eq_dd_az2}
\dot{\hat{\sigma}}_{z}(t) &=& -\gamma'(\hat{\sigma}_{z}(t)+1)+\hat{\zeta}_z \\ \nonumber
&-& 2 \int {\rm d} \omega \sin\left[k(L-l(t))\right] \,\left( g_{\omega} \hat{\sigma}_{+}(t) \hat{a}_{\omega}(t_0) \,e^{-i(\omega-\omega_a)t}+h.c.\right) \\ \nonumber
&-& 2 \int_{t_0}^{t} {\rm d} t'\,\int {\rm d} \omega \,{|g_{\omega}|}^2 \,\sin\left[k(L-l(t))\right] \,\sin[k(L-l(t'))]\\ \nonumber
&\times&\left(e^{-i(\omega-\omega_a)(t-t')} \hat{\sigma}_{+}(t)\hat{\sigma}_{-}(t')+h.c.\right).
\end{eqnarray}
\subsection{Characteristics of the quantum memory setup }
We will now focus in the quantum memory application of the considered setup and qualitatively discuss some characteristics the system should meet. One can thus make further assumptions which in turn enables to simplify the above equations.
Let's denote a round-trip time of the light between the atom and the mirror as $\tau = 2L/c$. In the ideal case we wish to absorb a photon by the atom, where the maximum coupling reaches $2\gamma_0$, as discussed earlier in this section, and indicates a relevant timescale (lower limit) for the photon duration. To prevent losses due to spontaneous emission during the write process, we thus require that (i) $\gamma_0 \tau \ll 1$ (Markov approximation). Furthermore, the coupling can be tuned between its maximal and minimal value by changing the atom-mirror distance on the order of the wavelength $\lambda$, thus changing the position of the atom between nodes and antinodes at will. We thus assume that (ii) $l(t) \approx \lambda$. Typically, $c \tau$ can be of the order of many wavelengths, so $\tau \gg l(t)/c$. With these arguments, we neglect the change in the operators on time scales smaller or equal to $\tau$, so that $\hat{\sigma}(t \pm \tau) \approx \hat{\sigma}(t \pm l(t)/c) \approx \hat{\sigma}(t)$. On the other hand, one must keep such dependence in all phases present in the equations in order to preserve the interferences. Then the atomic operators evolve as
\begin{eqnarray}
\label{eq dd_am3}
\dot{\hat{\sigma}}_{-}(t) &=& -\gamma(t) \,\hat{\sigma}_{-}(t) +\hat{\zeta}_- \\ \nonumber
&+& \hat{\sigma}_{z}(t)\int {\rm d} \omega g_{\omega}\,\sin\left[k(L-l(t))\right]\,e^{-i(\omega-\omega_a)t}\, \hat{a}_{\omega}(t_0),
\end{eqnarray}
\vskip -0.5cm
\begin{eqnarray}
\label{eq dd_az3}
\dot{\hat{\sigma}}_{z}(t) &=& - \gamma^z(t) \, \left(\hat{\sigma}_{z}(t)+1\right)+\hat{\zeta}_z \\ \nonumber
&-& 2 \int {\rm d} \omega \sin\left[k(L-l(t))\right] \,\left( g_{\omega} \hat{\sigma}_{+}(t) \hat{a}_{\omega}(t_0) \,e^{-i(\omega-\omega_a)t}+h.c.\right).
\end{eqnarray}
The time-dependent decay rates $\gamma(t)$ and $\gamma^z(t)$ are functions of the motion of the mirror $l(t)$
\vskip -0.5cm
\begin{equation}
\label{eq_gamma_t}
\gamma(t) = \frac{\gamma'}{2}+\frac{\gamma_p}{2}\, \left(1-e^{i\omega_a\left(\tau-\frac{2 l(t)}{c}\right)}\right),
\end{equation}
\vskip -0.5cm
\begin{eqnarray}
\label{eq_gamma_z_t}
\gamma^z(t) = \gamma'+ \gamma_p \, \left(1-\cos\left[\omega_a\left(\tau-\frac{2 l(t)}{c}\right)\right] \right)=2 Re [\gamma(t)],
\end{eqnarray}
\noindent where $\gamma_p$ is the decay into the pulse mode, which makes up the standard free space decay rate $\gamma_0$ together with the decay into the environment (the non-pulse mode) $\gamma'$, such that $\gamma' + \gamma_p = \gamma_0$. Using the Weisskopf-Wigner theory \cite[p. 207]{Scully_1997}, the explicit formula of $\gamma_p$ is given by $\gamma_p = \pi {|g_{\omega_a}|}^2$. We would like to note that in the derivation of the equations of motion \eqs{\ref{eq dd_am3}-\ref{eq dd_az3}}, various contributions to the level shifts are omitted (Lamb shift, Van der Waals and Casimir-Polder shifts). The reason is that for a typical atom-mirror distance $L \gg \lambda$, these level shifts are either negligible or constant \cite{Hetet_2010}. The only relevant dynamical level shift, which is the imaginary part of $\gamma(t)$ Eq. {\ref{eq_gamma_z_t}} is included.
With the general equations for the atomic operators \eqs{\ref{eq dd_am3}-\ref{eq dd_az3}} and the electric field operator, discussed more in detail in Appendix B \eqs{\ref{eq E+ scatt}--\ref{eq E2 scatt}}, it is now possible to study the dynamics of absorption, storage and retrieval of a single-photon wave packet. Since the absorption medium is a two-level system, we will consider in the following the storage process only of a single photon in Fock state \cite[p. 243]{Loudon_2000,Wang_2011}. The single photon Fock state pulse is defined as
\begin{equation}
\ket{1_{p}} = \int {\rm d} \omega \,f_p(\omega) \hat{a}^\dagger_{\omega} \ket{0} = \int {\rm d} t \,\xi_{p}(t) \hat{a}^\dagger_{t} \ket{0},
\end{equation}
\noindent where $f_p(\omega)$ is the spectral distribution function and $\xi_{p}(t)$ is the temporal shape of the wave packet, which are related by Fourier transform
\begin{equation}
\xi_{p}(t) =\frac{1}{\sqrt{ 2 \pi}} \int {\rm d} \omega \,f_p(\omega)\,e^{-i(\omega-\omega_0)t}.
\end{equation}
In the following, we use $p={\it {in},\it {out}}$ in order to label the input and output pulse waveform $\xi_{p}(t)$. Moreover, all the other considered quantities are labeled by {\it w} and {\it r} for the write and read process, respectively.
\subsection{Write process: Absorption}
\label{sec_write}
During the write process, we wish to efficiently absorb the incoming photon and thus maximize the probability $P$ that the atom gets excited, where ideally $P= 1$. Considering an incoming photon which is nonzero only between times $t_w$ and $t_w^0$, which are the start and end time of the write process, the write efficiency is defined as
\begin{equation}
\label{eq eta_w def}
\eta_w = \frac{P(t_w^0)}{\int_{t_w}^{t_w^0} {\rm d} t\, {|\xi_{in}(t)|}^2}.
\end{equation}
In the case of a single photon pulse, which satisfies the normalization condition $\int_{t_w}^{t_w^0} {\rm d} t\, {|\xi_{in}(t)|}^2=1$, the write efficiency is then simply $\eta_w = P(t_w^0)$. The excitation probability can be calculated using its definition
\begin{equation}
P(t) = \frac{1}{2}\,\Big( 1 + \bra{\psi(t_w)}\hat{\sigma}_{z}(t)\ket{\psi(t_w)} \Big),
\end{equation}
\noindent where $\ket{\psi(t_w)}=\ket{g,1_{in},0_e}$ is the initial state of the total system, with the atom being in its ground state, an incident single photon in Fock state and the environment is in the vacuum state.
So far we have included the environmental decay channel described by the decay rate $\gamma'$ and the related noise operators $\hat{\zeta}$. One important point is that when considering the initial state of the environment to be the vacuum state, the noise operators do not come into play, since $\bra{\psi(t_w)} \hat{\zeta} \ket{\psi(t_w)}=0$ (see also the discussion in \cite{Wang_2011}). Although it is very challenging to achieve experimentally, in the following we assume that all modes of the field radiated by the atom to the mirror half-space (i.e. to the left of the atom in Figure {\ref{fig setup}}) are covered by the mirror. This implies $\gamma'=0$, $\gamma_p = \gamma_0$. It also enables us to separate the effect of the time-dependent coupling $\gamma(t)$ from the effect of the decay to the environment. It is then clear from Eq. {\ref{eq_gamma_z_t}} that the time-dependent decay rate $\gamma^z(t)$ changes between $[0,2\gamma_0]$ depending on the position of the mirror.
The set of coupled differential equations \eqs{\ref{eq dd_am3},\ref{eq dd_az3}} for the atomic operators gives the absorption probability ({see Appendix A for details})
\begin{equation}
\label{eq Pe write}
P(t_w^0) = {\left|e^{-\Gamma_w{\left(t_w^0\right)}}\int_{t_w}^{t_w^0} {\rm d} t \,e^{\Gamma_w{(t)}} g_w(t) \,\xi_{in}(t) \right|}^2,
\end{equation}
where we define
\begin{equation}
\Gamma_w{(t)}=\int_{t_w}^{t} {\rm d} t' \gamma_w{(t')},
\end{equation}
\noindent with $\gamma_w{(t)}$ given by Eq. {\ref{eq_gamma_t}} and the subscript {\it w} indicates the write process in order to distinguish it from the read process which has in principle different decay function $\gamma_r{(t)}$ . The effective time-dependent coupling strength reads
\begin{equation}
g_w(t)= \sqrt{2 \gamma_0} \, \sin \left[\omega_a\left(\frac{\tau}{2}-\frac{l(t)}{c}\right)\right]=\sqrt{\gamma_w^z(t)}.
\end{equation}
The goal is now to find the time-dependent $\gamma_w^z(t)$ that maximizes the write efficiency for a given input field $\xi_{in}(t)$, which can be done using Lagrange multiplier optimization \cite[p. 169]{Riley_2006},
\begin{equation}
\label{eq lag}
\frac{\delta}{\delta \xi^*_{in}(t)}\left[P(t_w^0)+\lambda \left(\int_{t_w}^{t_w^0} {\rm d} t\, {|\xi_{in}(t)|}^2-1\right)\right]=0
\end{equation}
\noindent where $\lambda$ is the Lagrange multiplier.
This results in the optimized write efficiency
\begin{equation}
\label{eq etaw}
\eta_w = 1- e^{-\Gamma_w^z(t_w^0)},
\end{equation}
\noindent with $\Gamma_w^z{(t)}=\int_{t_w}^{t} {\rm d} t' \gamma_w^z{(t')}$, and the time-dependent decay rate satisfying
\begin{equation}
\label{eq_gamma_zw}
\gamma_w^z(t)=
\left\{
\begin{array}
{r@{\quad:\quad}l}
\frac{\eta_w\,{|\xi_{in}(t)|}^2}{(1-\eta_w)+\eta_w \int_{t_w}^{t_w^0} {\rm d} t'\, {|\xi_{in}(t')|}^2} & \gamma_w^z(t) \leq 2 \gamma_0;\\
2 \gamma_0 & \gamma_w^z(t) \geq 2 \gamma_0,
\end{array}
\right.
\end{equation}
\noindent where we have to account for the physical limitation of the system, $ 0 \leq \gamma_w^z(t) \leq 2\gamma_0$.
After the absorption, the single photon is stored as the excitation of the atom for a time period $T$. During this period, the static mirror position is such that the atom sits at the node, i.e. $\gamma_w^z(t)=0$, so that the atom remains in its excited state, which implies that $P(t_w^0 \leq t \leq t_r^0) = P(t_w^0)$ during the storage period.
\subsection{Read process: Re-emission}
\label{sec_read}
For an on-demand readout of the stored single photon pulse, the atom-light interaction is turned on again at the starting time of the readout process $t_r^0 = t_w + T$. As discussed above, we consider no losses during the storage process, so that $P(t_r^0)=P(t_w^0)=\eta_w$. In analogy to the write efficiency, we define the efficiency of the readout process ending at time $t_r$ as
\begin{equation}
\label{eq eta_r def}
\eta_r = \frac{\int_{t_r^0}^{t_r} {\rm d} t\, {|\xi_{out}(t)|}^2}{P(t_r^0)}.
\end{equation}
The temporal shape of the outgoing pulse $\xi_{out}(t)$ can be derived from the electric field operators \eqs{\ref{eq E+ scatt}--\ref{eq_xiout_def}} in Appendix B as
\begin{eqnarray}
\label{eq_xiout1}
\xi_{out}(z,t) &=& \sqrt{\frac{2}{\pi}}\, \frac{1}{A(\omega_a)} \bra{\psi_0}\hat{E}_{out}^+(z,t)\ket{\psi(t_r^0)} \\ \nonumber
&=& i \sqrt{\frac{2}{\gamma_0}} \, e^{-i \omega_a (t-z/c+\tau/2)}\,\gamma_r(t) \bra{\psi_0} \hat{\sigma}_{-}(t-z/c) \ket{\psi(t_r^0)},
\end{eqnarray}
\noindent with $\ket{\psi_0}=\ket{g,0_{in},0_e}$ and $\ket{\psi(t_r^0)}=\ket{e,0_{in},0_e}$.
The evolution of the atomic operators can be also found using \eqs{\ref{eq dd_am3}-\ref{eq dd_az3}}
\begin{equation}
\label{eq_am_r1}
\bra{\psi_0} \hat{\sigma}_{-}(t-z/c) \ket{\psi(t_r^0)} = \sqrt{P(t_r^0)} \,e^{-\Gamma_r(t)}.
\end{equation}
Since we are interested in the output pulse at certain position $z \geq L$, the temporal shape of the output pulse $\xi_{out}(t)$ reads
\begin{equation}
\label{eq_xiout2}
\xi_{out}(t) = \xi_{out}(z,t)\big |_{z=D \geq L} = i \sqrt{\frac{2\,P(t_r^0)}{\gamma_0}} \, e^{-i \omega_a (t-D/c+\tau/2)}\,\gamma_r(t)\,e^{-\Gamma_r(t)},
\end{equation}
\noindent which implies that the temporal shape of the output pulse can be adjusted by controlling the time-dependent read decay rate
\begin{equation}
\gamma^z_r(t)=\frac{{|\xi_{out}(t)|}^2}{\eta_w - \int_{t_r^0}^{t} d t'\, {|\xi_{out}(t')|}^2},
\end{equation}
which is again subjected to the constraint that $ 0 \leq \gamma_r^z(t) \leq 2\gamma_0$.
Plugging Eq. {\ref{eq_xiout2}} into Eq. {\ref{eq eta_r def}} one finds the expression for the read efficiency
\begin{equation}
\label{eq etar}
\eta_r = 1- e^{-\Gamma_r^z(t_r^0)},
\end{equation}
\noindent with $\Gamma_r^z{(t)}=\int_{t_r^0}^{t} {\rm d} t' \gamma_r^z{(t')}$.
The total quantum memory efficiency is given by
\begin{equation}
\eta=\eta_w\,\eta_r=(1- e^{-\Gamma_w^z(t_w^0)})\,(1- e^{-\Gamma_r^z(t_r^0)}).
\end{equation}
So far we have derived an expression for the efficiency of the readout process as a function of a time-dependent readout decay rate $\gamma^z_r (t)$. We should however emphasize a simple reflection that, an atom in the excited state with a nonzero coupling to the field will necessarily decay. Typically, for a constant $\gamma_r$, the decay will be exponential with $\eta_r$ approaching 1 already for times of $1/\gamma_r$. The readout can be thus made simply by ``waiting".
In the following, we would rather require that the quantum memory device yields the maximum fidelity $F =1$.
The memory fidelity is expressed in terms of the outgoing pulse's projection on the input pulse as
\begin{eqnarray}
\label{eq F def}
F = \left| \bra{1_{in}} 1_{out} \rangle \right|^2 = \frac{{\left|\int d t \, \xi^*_{in}(t)\xi_{out}(t)\right|}^2}{\int d t \, {\left|\xi_{in}(t)\right|}^2 \cdot \int d t \,{\left|\xi_{out}(t)\right|}^2}.
\end{eqnarray}
Obviously, the ideal fidelity is achieved when the output pulse has the same shape as the input pulse $\xi_{out}(t)= \sqrt{\eta} \,\xi_{in}(t-T)$, which can be achieved by changing the read decay rate in the following way,
\begin{equation}
\label{eq_gamma_zr}
\gamma_r^z(t)=
\left\{
\begin{array}
{r@{\quad:\quad}l}
\frac{\eta_r\, {|\xi_{in}(t+(t_r^0-t_w^0))|}^2}{1 - \eta_r \,\int_{t_r^0}^{t} d t'\, {|\xi_{in}(t'+(t_r^0-t_w^0))|}^2} & \gamma_r^z(t) \leq 2 \gamma_0;\\
2 \gamma_0 & \gamma_r^z(t) \geq 2 \gamma_0.
\end{array}
\right.
\end{equation}
Due to the similarity of the underlying physics, we would like to note that the expressions for read and write efficiency \eqs{\ref{eq etaw},\ref{eq etar}} are analogous to those in Ref. \cite{Green_2011}.
\section{Simulations and possible implementations}
\label{sec Sim_Imp}
\subsection{Simulation with time-bin qubit}
\label{sec Simulation}
With the help of \eqs{\ref{eq_gamma_zw},\ref{eq_gamma_zr}}, we can now study the performance of the quantum memory as a function of the input light field. In the following, we consider a specific case of a normalized Gaussian-shaped time-bin single photon pulse described as
\begin{equation}
\label{eq_tb}
\xi_{in}(t) = \alpha\,e^{-\frac{{(t-t_{1})}^2 \sigma^2}{2}} + \beta\, e^{i \phi}\,e^{-\frac{{(t-t_{2})}^2 \sigma^2}{2}},
\end{equation}
\begin{figure}[h!]
\begin{minipage}{1 \linewidth}
\includegraphics[scale=0.32]{tb_small_xir.pdf}
\vspace{0.0cm}
\end{minipage}
\begin{minipage}{1 \linewidth}
\includegraphics[scale=0.32]{tb_large_xir.pdf}
\vspace{0.0cm}
\end{minipage}
\caption{(Color online) Storage ($t<0$) and retrieval ($t>0$) of a Gaussian-shaped time-bin single photon pulse for different values of bandwidth: (a) $\sigma =0.2 \gamma_0$ ; (b) $\sigma=5 \gamma_0$. The input intensity (dashed black line) and the output intensity (solid black line) of the pulse is shown in (i), with the input intensity normalized to amplitude 1. The required optimum write and read decay rate $\gamma^z_w(t)$ and $\gamma^z_r(t)$ is shown by dashed and solid red line in (ii), respectively. It can be seen that for the smaller bandwidth, case (a), $ {\it Max} [\gamma^z_{w,r}(t)] < 2 \gamma_0$ and the efficiency is close to $1$; on the other hand, for the larger bandwidth, case (b), where $\gamma$ has to be truncated at $2 \gamma_0$, the efficiency is less than $1$. }
\label{fig inout}
\end{figure}
\noindent where the real coefficients $\alpha,\,\beta$ satisfy $\alpha^2 + \beta^2 =1$, $t_{2}-t_{1}$ is the relative time delay, $\phi$ is the relative phase between the two time bins and the bandwidth $\sigma$ is assumed the same for each time bin. The performance of the quantum memory is studied for different bandwidths $\sigma$ of the pulse with $\alpha=\beta$. In Figure {\ref{fig inout}}, two particular situations are considered, one with photon bandwidth smaller and the other one with photon bandwidth larger than the double of the atomic decay rate $2 \gamma_0$. In Figure {\ref{fig inout}} (a), we set $\sigma =0.2 \gamma_0$. In this case, the quantum memory efficiency reaches its maximal value, $\eta =1$: the amplitude of the output pulse (solid black line) is the same as the input pulse (dashed black line) as can be seen from Figure {\ref{fig inout}} (a)(i). On the other hand, Figure {\ref{fig inout}} (b) with $\sigma =5 \gamma_0$ shows a decrease of the efficiency. The optimized decay rates $\gamma^z_w(t)$ and $\gamma^z_r(t)$ are represented by dashed and solid red lines respectively. The shapes of the optimum coupling decay rates are given by \eqs{\ref{eq_gamma_zw},\ref{eq_gamma_zr}} and might be qualitatively understood as follows. For write efficiencies $\eta_w \approx 1$, the write decay rate $\gamma^z_w(t)$ is proportional to the intensity divided by the time integral of the intensity. This ratio can be high at the beginning of the write process (first time bin), when the denominator is small, but gets significantly smaller for the second time bin. Similar argument holds for the read coupling decay. It is possible to plot the motion of the mirror $l(t)$ instead of the coupling decay rate $\gamma^z(t)$ (see Eq. {\ref{eq_gamma_z_t}}). In the example presented in Figure {\ref{fig inout}}, the motion of the mirror is similar to the coupling decay rate with the $l(t)$ ranging from 0 to $\lambda/4$ (corresponding to $2\gamma_0$ for the decay rate) and we do not plot it explicitly. Finally, one can see that for the photon bandwidth larger than the cutoff frequency of the system $2\gamma_0$, the optimum decay rates $\gamma^z_{w,r}(t)$ exceed this cutoff and are thus truncated at $2\gamma_0$. This results in the decrease of the storage efficiency, as shown in Figure {\ref{fig inout}} (b)(i). The storage efficiency as a function of the ratio between the photon bandwidth and the atomic decay rate is shown in Figure {\ref{fig eta}}. The efficiency starts to decrease for $\sigma/\gamma_0 \approx 0.85$ which corresponds to $ {\it FWHM } =2 \sqrt{2{\rm Log}2} \,\sigma = 2 \gamma_0$, as expected.
\begin{figure}[h!]
\includegraphics[scale=0.3]{tb_eff.pdf}
\caption{ Total efficiency of the quantum memory device as a function of the bandwidth $\sigma$ of the input pulse. The curve was obtained with Gaussian-shaped time-bin single photon wave packet Eq. {\ref{eq_tb}} for ideal fidelity $F=1$. The inset is a zoom of the region of $\sigma/\gamma_0$ between $0$ and $2$.}
\label{fig eta}
\end{figure}
\subsection{Implementations}
\label{sec Implementations}
We will now discuss possible implementations of our protocol. The described quantum memory device requires a single two-level system with a tunable distance to the mirror and a strong coupling to the light field. Strongly coupled two-level systems can be implemented using optical setups with ions and atoms \cite{Eschner_2001, Tey_2008}, quantum dots \cite{zhang_2008}, superconducing qubits in circuit QED configuration \cite{Devoret_2004, You_2011} or atoms coupled to surface plasmons on conducting nanowires \cite{Chang_2007} or to tapered optical nanofibers \cite{Nayak_2011}.
As for the quantum optical implementations, there is a variety of atoms and ions used in trapping experiments, typical examples being e.g. a $^{138}{\rm Ba}^+$ ion in a Paul trap \cite{Eschner_2001, Slodicka_2010, Hetet_2011} or $^{87}{\rm Rb}$ atom in a FORT trap \cite{Tey_2008}. In the case of ${\rm Ba}^+$ ions, the typical two-level transition is between the ground state $\ket{6S_{1/2},m_F=1/2}$ and excited state $\ket{6P_{1/2},m_F=-1/2}$ at $\lambda=493$ nm with a linewidth $\gamma_0 = 15$ MHz \cite{Slodicka_2010}. For this system, as experimental realization with half-cavity an tunable atom-mirror has been reported \cite{Dubin_2007} (an analogous experimental setup with quantum dot has been also realized \cite{zhang_2008} ). This, together with an atom-mirror distance $L$ of order of centimeters, meets very well the assumption required for quantum memory: $\gamma_0\tau \ll 1$. On the other hand, the durations of incoming photon of the order up to $1/\gamma_0$ require the motion of the mirror at the same time scale, which might be hard to achieve by a mechanical motion. One possible solution is to use a long-lived quadrupole transition (for which the lifetime can be seconds (e.g. Ca$^{+}$ or Ba$^{+}$)) which would allow for slower mechanical motion of the mirror achievable with current technology. Another possibility is to move the atom itself, which can be done very fast in Dipole or Paul traps. The drawback of this approach is that the atom would get slightly out of the focus of the mirror, reducing thus the maximum achievable coupling decay rate \cite{Hetet_2010}.
It might be also be possible to use an EOM in the integrated setup to modulate the optical path-length \cite{Chang_2007, Nayak_2011}.
The spatial overlap of the incident field and the atomic dipole pattern needs to be taken into account in realistic systems, as discussed elsewhere in more detail for hemispherical mirror \cite{Hetet_2010} and for parabolic mirror \cite{Sondermann_2008, Stobinska_2009}. The consequence of imperfect spatial overlap is the decay into the environment $\gamma'$ which would reduce the write efficiency as well as, and more importantly, the storage process (since the storage time $T$ is often required to be much larger than the photon duration, the population of the excited state $\propto {\rm exp}(-\gamma' T)$ is more affected during the storage, because ${\rm exp}(-\gamma' T) \ll {\rm exp}(-\gamma't_p)$, where $t_p$ is the pulse duration). Obviously, the quantum memory scheme works only for single photon Fock states, which are available experimentally \cite{Kuhn_2002}. Finally, we would like to mention that the quantum memory works also for the polarization qubits. In this case the required level scheme is a V configuration, standardly available for typical atoms used in the experiments.
The proposed quantum memory device can be also implemented in the fast growing domain of circuit QED, where the effective two-level system can be realized by different kinds of superconducting qubits \cite{Devoret_2004, You_2011}. Typical resonant frequencies of a superconducting qubit lay in the microwave region of order of 1-10 GHz with population decay rates of order of 1-10 MHz \cite{Wallraff_2004, Houck_2007, Abdumalikov_2010}. Generation of various photonic states, including a single photon Fock state, was demonstrated in several experiments \cite{Houck_2007, Hofheinz_2008, Hofheinz_2009} laying thus the ground for potential realization of the presented quantum memory scheme. The configuration of superconducting qubits coupled to a transmission line resonator has the beauty of well defined one-dimensional (1D) mode and perfect spatial overlap, which results in strong atom-light interaction. Moreover, an open transmission line with one side terminated by a SQUID operated with a variable magnetic flux, acts as a mirror with a tunable qubit-mirror distance. This was realized recently in the remarkable demonstration of dynamical Casimir effect by Wilson {\it et al.} \cite{Wilson_2011}, with oscillation frequency of the SQUID mirror of 11 GHz. Currently, schemes and proposals directly linked to the quantum memory applications are actively investigated both theoretically \cite{Marcos_2010} and experimentally \cite{Schuster_2010, Wu_2010, Mariantoni_2011_nphy, Mariantoni_2011_sci}. In one of the realized experiments, a superconducting qubit with a large decoherence rate (order of MHz) was coupled to a transition with a long coherence time (up to 2 ms) in a NV center in a diamond \cite{Kubo_2010}. This technique can be applied also to our proposal to achieve long storage time for the microwave photons.
\section{Conclusion}
\label{sec Conclusion}
In conclusion, we showed by a fully quantized calculation, that a single photon Fock state pulse with various temporal shapes can be efficiently stored and retrieved from a quantum memory device consisting of a single two-level atom in a half cavity. The principle is that the time-dependent atomic decay rate can be dynamically tuned between zero and the maximum $2 \gamma_0$ by changing the distance between the atom and the mirror. The cutoff frequency of the system, given by double of the free space decay rate of the atom, imposes the limits on the input photon bandwith for which the photon can be efficiently stored. We analyzed the dependence of the storage efficiency as a function of the photon bandwidth. Finally, we discussed possible implementations of the proposed quantum memory scheme, such as single atoms/ions in a half cavity or a superconducting qubit coupled to a 1D transmission line terminated by a SQUID.
~\\
\section{Acknowledgements}
We would like to thank Colin Teo, Jing Yan Haw and Luk\'a\v{s} Slodi\v{c}ka for useful discussions. This work was supported by the National Research Foundation and the Ministry of Education, Singapore. G. H. acknowledges support by a Marie Curie Intra-European Action of the European Union.
\bibliographystyle{prsty}
|
\section{Introduction}
\gerda\ \cite{gerda} is a low-background experiment searching for the
neutrinoless double beta decay of $^{76}$Ge, using an array of bare high-purity
germanium (HPGe) detectors isotopically enriched in $^{76}$Ge.
The detector array is operated directly in ultra radio-pure liquid argon, allowing a
substantial background reduction at the $Q_{\beta\beta}$-value of $^{76}$Ge with
respect to the previous experiments~\cite{hm}.
In the present phase (Phase\,I) eight enriched coaxial detectors are being used,
totaling approximately 15 kg of $^{76}$Ge.
In Phase\,II, about 30 new custom-made enriched BEGe
detectors~\cite{bege} will be deployed (additional $\sim$20\,kg of $^{76}$Ge).
The experiment is located in the underground Laboratori Nazionali del Gran Sasso
of the INFN (Italy).
The background suppression in \gerda\ is achieved by the specific innovative
design (namely, detectors operated naked in a cryogenic liquid) and by a strict
material selection. In addition, part of the remaining background
events can be identified by an off-line analysis of the HPGe detector signals,
i.e. detector anti-coincidence and pulse shape discrimination
techniques~\cite{psd}.
For this purpose, a software framework (\gelatio) for advanced digital signal
processing and analysis has been recently developed~\cite{gelatio}.
It is implemented in C++ and is based on the \textsc{MGDO} library~\cite{mgdo}.
The framework is designed to support a multi-channel data processing and a
modular analysis of digital signals.
Signals are analyzed by using chains of modules completely customizable by
the user.
Each module handles a precise and self-consistent task of the signal
processing and is implemented as a dedicate C++ class.
The output of the modules, which is either a scalar parameter (e.g. the
amplitude of the signals) or a shaped trace, can be used as input for other
modules and/or stored to disk.
The framework was used for the reference analysis of the data acquired in
the \gerda\ commissioning phase (from June 2010 to October 2011), when up to
seven HPGe detectors have been operated simultaneously.
The commissioning data were used as a benchmark to validate \gelatio\ against
other independent analysis codes and to prove its suitability for
the use in the \gerda\ Phase\,I.
This paper describes the basic off-line analysis of the \gerda\ data performed
with the \gelatio\ framework.
In section~\ref{dsp} we present the flow of the signal processing and analysis
along the module chains and the shaping algorithms.
Then, in section~\ref{filter}, we discuss the identification of non-physical
events or of signals not properly processed along the analysis pipeline. Also, the
monitoring of the data quality will be described.
Finally, summary and conclusions are presented in section~\ref{last}.
\section{Signal processing flow}\label{dsp}
The charge pulses from the HPGe detectors
operated in \gerda\ are digitized by
14-bit flash-ADCs~\cite{daq} (FADC) running at 100\,MHz sampling rate.
For each event, the FADC computes in run-time two traces that are eventually
written to disk.
The first trace is sampled at 100\,MHz and is 4\,$\mu$s long
(high-frequency-short trace).
It includes the signal leading edge and it is used for identifying
background events through pulse shape discrimination techniques.
The second trace has a sampling frequency of 25\,MHz and is
160\,$\mu$s long (low-frequency-long trace).
It is used for those operations, as energy reconstruction, which involve
the integration of the signal.
The two traces are processed along different chains of \gelatio\ modules, as shown in
\figurename~\ref{calflow}.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.9\textwidth]{fig/gerdaFlow4}
\end{center}
\caption{\small Flow chart of the signal processing. The two traces saved by
the digitizer are processed along two different chains of \gelatio\ analysis modules.
The
low-frequency-long (LFL) trace is used for reconstructing the energy, the
trigger and the rise time. The high-frequency-short (HFS) trace is used for
pulse shape analysis.}
\label{calflow}
\end{figure}
The first module is GEMDTop. It takes care of extracting the traces from the
input file and making them available to the other modules. It also checks, and
possibly changes, the pulse polarity in order to always have positive-polarity
pulses. The output traces are the starting point for two chains: the
low-frequency-long trace is processed along Chain1 while
the high-frequency-short trace along Chain2.
Chain1 starts with GEMDBaseline.
This module analyzes the baseline of the signal by computing the average
value, the rms and the linear slope in the pre-trigger region.
In addition, the module performs a baseline restoration --- a subtraction of the
average baseline value to the trace --- and provides the new signal
to the other modules:
\begin{itemize}
\item GEMDTrigger. The module implements a leading-edge discriminator with
threshold defined dynamically as three times the rms of the signal
baseline.
After the trigger, the signal has to remain above threshold for
at least 40\,$\mu$s,
otherwise the trigger is rejected.
Before searching for the trigger, the pulse is integrated using a
160\,ns moving average filter in order to reduce the high-frequency noise
\item GEMDFTTrigger.
The module applies to the input signal a 1.5\,$\mu$s moving
differentiation filter and a 1\,$\mu$s moving average filter for noise
reduction (see
\figurename~\ref{shaping1}).
The resulting trace has a peak for each sharp variation of the signal
(such as the leading edge of a pulse)
and is analyzed by a leading-edge discriminator.
The peak width is similar to the size of the moving differentiation and
was chosen to maximize the pile-up identification efficiency and to
avoid the mis-identification of highly-multiple-site events.
The number and the position of the peaks are estimated by
applying a leading-edge discriminator, whose threshold is
four times the rms of the baseline.
After this condition is met, the signal must remain above the threshold for
at least 1\,$\mu$s.
While GEMDTrigger is tuned to determine the trigger position with high
precision and stability (it requires that the signal remains above the
threshold for 40\,$\mu$s),
this module is important to identify
events with multiple physical signals occurring within the same trace.
\item GEMDEnergyGauss. The module reconstructs the event energy using an
approximate Gaussian filter~\cite{dspguide}.
The pulse is differentiated by a moving differentiation filter
and then integrated 15 times by a moving average filter to achieve an
approximated Gaussian shape\footnote
Historically, the energy reconstruction filters for $\gamma$-ray spectroscopy
also perform a deconvolution of the exponential function
which is folded in the signal by the charge sensitive pre-amplifier.
However, the approach suggested in this paper was found to provide better
results on the \gerda\ data with respect to the usual filters.
}.
The energy information is eventually stored
in the maximum amplitude of the quasi-Gaussian pulse.
The width of the moving filters has been set to 10\,$\mu$s in order to minimize
losses due to ballistic effects.
The intermediate steps of the shaping are shown in
\figurename~\ref{shaping2}.
\begin{figure}[tbp]
\begin{minipage}[t]{0.48\textwidth}
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/ftt2}
\caption{\small
Digital signal processing performed by GEMDFTTrigger.
The incoming signal~(top trace) is differentiated (middle) and then
integrated.
Each sharp variation of the incoming pulse creates a peak in the output trace
(bottom trace).}
\label{shaping1}
\end{center}
\end{minipage}
\hspace{0.02\textwidth}
\begin{minipage}[t]{0.48\textwidth}
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/gauss}
\caption{\small
Digital signal processing performed by GEMDEnergyGauss.
The incoming signal (top trace) is differentiated
(second trace) and then integrated several times (following traces)
by a moving average filter.}
\label{shaping2}
\end{center}
\end{minipage}
\end{figure}
\item GEMDRiseTime. The module computes the rise time between 10\% and 90\% of
the maximum amplitude of the pulse.
The maximum amplitude is computed as the difference between the maximum of the
pulse and the average baseline value.
Then, the first samples below the 10\% and 90\% of the maximum amplitude
are found by moving backwards from the position of the maximum.
\end{itemize}
The second chain is used to evaluate parameters relevant for pulse shape
discrimination techniques and it will be better defined during the future data
taking.
The chain presently includes only one module, GEMDCurrentPSA, which computes the
current pulse as the derivative of the signal and then extracts the basic
features of the current peak, like rise time, width and area.
\section{Data selection and monitoring}\label{filter}
In the \gerda\ data sets there are two main classes of signals that have to be
identified and tagged: 1) signals corrupted or produced by non-physical events, i.e.
discharges, cross-talk, pick-up noise; 2) signals which are not properly
processed along the analysis pipeline, as pile-ups and accidental coincidences.
The first class includes signals with
anomalous shape, wrong polarity, extremely short/long rise time or exceeding
the dynamic range of the FADC (see \figurename~\ref{filter1}).
To identify these events a sequence of cuts based on four parameters is applied.
The first parameters are the trigger position computed by GEMDTrigger and the
time position of the maximum amplitude of the Gaussian pulses (maxAmpTime)
computed by GEMDEnergyGauss.
If the signal has a leading edge at the proper position,
followed by an exponential decay tail due to the charge sensitive pre-amplifier,
then the trigger has to be reconstructed roughly in the center of the trace and
maxAmpTime has to be in a well-defined range.
The third parameter is the 10-90\% rise time which can be used to identify
signals that are extremely fast or slow, and hence inconsistent with well-behaved
physical events. Finally, signals that saturate the dynamic
range of the FADC are identified by scanning the individual
samples of the traces.
The second class includes signals generated by the superimposition of multiple physical
pulses, or having the leading edge not aligned with the center of the
trace (see \figurename~\ref{filter2}).
\begin{figure}[t]
\begin{minipage}[t]{0.48\textwidth}
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/filter1}
\caption{\small
Illustrative traces generated by non-physical events. Note that these
signals do not have the typical exponential decay tail after the leading edge.
}
\label{filter1}
\end{center}
\end{minipage}\hspace{0.02\textwidth}
\begin{minipage}[t]{0.48\textwidth}
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/filter2}
\caption{\small
Example of traces generated by pile-up signals (top-middle trace) and
accidental coincidences (bottom).
}
\label{filter2}
\end{center}
\end{minipage}
\end{figure}
These signals can be identified using the baseline slope computed by
GEMDBaseline, the number of pulses provided by GEMDFTTrigger and the trigger
position yielded by GEMDTrigger.
Their amount is proportional to the event rate and can reach up to 15\% in
the calibration data sets, while it is usually negligible in the physics data sets.
Therefore, the identification of these signals is a critical issue to
derive from the calibration data a sample of events which is as similar
as possible to the physics run data.
Besides cuts for removing undesirable classes of signals,
there are also parameters which can be used to monitor the quality of the data taking
and the stability of the set-up, the most important being the average value
and the rms of the baseline. These parameters are sensitive to noise changes and to
gain variations in the read-out chain.
\figurename~\ref{profile} shows these parameters as a function of time for a 10-day
commissioning run.
\begin{figure}[t]
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/profile}
\end{center}
\caption{\small
Average value (top panel) and rms (bottom panel) of the signal baseline vs. time,
for a detector operated in \gerda\ during a 10-day run.
The bin content represents the mean value of the parameter and the error
bars are related to the width of the distribution. The bins are 2~hours wide.
}
\label{profile}
\end{figure}
The parameters are stable over the whole data taking, except for a few hours
during day 7. These instabilities can be correlated with hardware operations in
the set-up and the corresponding data can be removed by applying a cut on the
two parameters.
The results of the data selection performed according to the criteria
described above are shown in \figurename~\ref{spectrum}.
The figure shows the energy spectrum of a HPGe detector operated in
the \gerda\ set-up irradiated with a $^{228}$Th source.
The cuts remove efficiently bad signals and pile-up events, improving the
shape of the $\gamma$-ray peaks and the agreement with the standard
analytical functions used to model $\gamma$-line peaks.
\begin{figure}[tbp]
\begin{center}
\includegraphics[angle=270,width=\textwidth]{fig/spectrum}
\end{center}
\caption{\small
Reconstructed energy spectrum of an enriched detector deployed in the \gerda\
set-up irradiated with a $^{228}$Th source. The spectrum is shown before (blue
histogram) and after (red histogram) having applied the cuts described in
section~\ref{filter}. The cuts remove approximately 15\% of the total events
($\sim$10\% in the $\gamma$ lines). The bottom panels show the peak of the
2.614\,MeV $\gamma$-ray line and the analytical model used for the fit, in
linear and logarithmic scale. The tail on the left-hand side is due to pile-up
events and is visibly reduced by the cuts.
}
\label{spectrum}
\end{figure}
\section{Conclusions}\label{last}
The \gerda\ experiment is currently starting the data taking of Phase\,I.
The off-line analysis of the HPGe detector signals will be performed with the
\gelatio\ software framework, a tool specifically developed for \gerda, which
supports a multi-channel analysis and implements a signal processing based on a
modular approach.
A reference analysis pipeline has been defined and optimized.
The signal
processing is performed along chains of modules and includes the estimate
of the trigger position, of the amplitude and of several
basic pulse shape analysis parameters.
The digital filters have been improved and optimized during the \gerda\
commissioning phase and the shaping parameters have been tuned.
Also, a set of cuts has been defined to identify signals
induced by non-physical events or signals which are not properly processed.
In addition, a set of parameters was identified to monitor
the data quality and possibly to discriminate corrupted data.
The software and the digital filters have been validated during the \gerda\
commissioning and in several R\&D activities related to the experiment.
The new pipeline has been tested and it proved to be stable and ready to be used
for the reference analysis of \gerda\ Phase\,I data.
\ack
We would like to acknowledge our colleagues of the \gerda\ Collaboration,
especially D.~Budj\'{a}\v{s} and B.~Schwingenheuer, for many invaluable advices
concerning signal analysis and data quality control.
We want also to thank C. A. Ur, D. Bazzacco and T. Kihm for all the stimulating
discussions and suggestions concerning digital signal processing and the
algorithms used in $\gamma$-ray spectroscopy.
This work was supported in part by the Transregio Sonderforschungsbereich
SFB/TR27 ``Neutrinos and Beyond'' by the Deutsche Forschungsgemeinschaft and by
the Munich Cluster of Excellence ``Origin and Structure of the Universe''.
\section*{References}
|
\section{Introduction}
Amongst the axisymmetric constituents of the plasma flow in the Sun's
convection zone, the meridional circulation (MC), although being much
slower than the differential rotation, is gaining growing interest which
arises from its potential importance for the solar dynamo.
Certainly any advanced non-linear mean-field model of this dynamo
will have to include the MC
to cover the full interaction between
mean field and motion.
In particular however, flux-transport dynamo models
depend already on the {\em kinematic level} crucially on the MC as
it is the ingredient which allows to close the dynamo cycle.
At the same time it explains naturally both
the equatorward migration of active regions
and the poleward drift of weak poloidal field
elements during the solar cycle.
The MC in the Sun is accessible to a variety of techniques
out of which Doppler measurements
and helioseismological time-distance or ring diagram analyses
tend to agree in the surface speed
profiles with a peak of $\approx20$m/s
at a latitude $\approx35^{\circ}$. However, attempts to
measuring the MC
velocity using small magnetic structures as tracers of the flow
(\cite[Komm et al., 1993, Hathaway \& Rightmire, 2010]
{komm+etal_93,hatha+righ_10}) reveal systematic differences from
these results.
Compared to Doppler measurements (\cite[Ulrich 2010]{ulrich2010})
magnetic feature tracking (MFT) speeds
can be as much as $\approx 30$\% smaller.
This indicates that the magnetic field is {\it not completely frozen}
in the plasma, which is reasonable given the
assumed high turbulent magnetic diffusivity $\eta_{\rm T}$. Due to
diffusive spreading of the field, its
advection is affected by the depth
dependence of the MC velocity.
On top of advection and diffusion,
turbulent motions may be contributing by
an extra advection term due to turbulent magnetic pumping
or by the more complicated interaction with dynamo waves.
In a previous work (\cite[Guerrero et al., 2011]{guerr+etal_11}, hereafter
GRBD) we studied numerically the
kinematic evolution of small bipolar magnetic elements at
different latitudes in the northern hemisphere.
For simplicity we considered turbulent diffusion
and advection due to a predefined MC
as the only magnetic transport mechanisms. In a 1D
(i.e., surface) version of the model, the
northern spot acquires an effective velocity which is higher
than the flow speed, but the southern
one travels slower than the fluid.
However, on average,
or considering the center of the bipolar region, its
speed clearly fits the fluid speed.
This result is independent of the value of $\eta_{\rm T}$.
In the 2D model, on the other hand, we found that the MFT speed
differs from the fluid velocity provided
that the frozen-in condition is not satisfied.
Moreover, we demonstrated that the
relevant threshold value of $\eta_{\rm T}$
depends on the radial gradient of
the latitudinal velocity, $\partial_r |U_{\theta}|$,
assumed positive. The difference between MFT and flow speeds
increases with $\partial_r |U_{\theta}|$.
Thus, this simple model is able to explain the discrepancy
between the Doppler and MFT observations.
Modern local helioseismological techniques
allow also to infer the radial variation of the MC
speed. Unfortunately, time-distance and
ring-diagram analyses have not given consistent results.
The first method yields that at the first few
megameters beneath
the photosphere the latitudinal velocity decreases inwards,
i.e., a positive $\partial_r |U_{\theta}|$
(\cite[Zhao et al. 2011]{zhao+etal_11}).
By contrast, the results from the second method indicate a
slight inward {\em increase}, i.e., a negative gradient
albeit restricted to middle latitudes.
(\cite[Gonz\'{a}les Hern\'{a}ndez et al. 2006]{gh+etal_06}).
Such velocity profiles were not considered in GRBD and
the aim of this paper is to fill this gap as well as to
discuss other effects that may influence MFT speeds.
\section{MFT speed with a negative radial gradient of $\,|U_{\theta}|$}
The method followed here to simulate
the evolution of small bipolar magnetic structures is described
in detail in GRBD. We also use the same notation.
Briefly, we solve the induction equation
in axisymmetry for the azimuthal component of the vector potential,
$A_{\phi}$ in the variables $r$ and $\theta$. The prescribed meridional
flow fed into the model is given by eqs. (4-6) of GRBD
and has a single cell per hemisphere. In the present paper
$\eta_{\rm T}$ is assumed to be
constant across the domain. Bipolar regions
(with spot separation ~3$^{\circ}$ and depth $0.04R$) at
20 different latitudes from the pole to equator are
evolved over a time interval of two weeks. Their
velocity is computed as an average over this time span
through the differences in final and initial latitudes
(see eq. (8) of GRBD).
\begin{figure}
\centering
\vspace{-2mm}
\includegraphics[width=0.49\columnwidth]{uthrad}
\includegraphics[width=0.49\columnwidth]{pos_grad}
\caption{
a) Radial profile of $U_\theta(r,57^\circ)$ for different
values of $n$. Dashed line: radial profile
of $\gamma_{\theta}$, see eq. \eqref{eq:thp}.\; b) Corresponding
tracking velocities, $\mean{V}$,
for three cases: $\partial_r |U_{\theta}|>0$ (yellow/continuous) and
$\partial_r |U_{\theta}|<0$ with
(dashed) and without (red/dot-dashed)
pumping. $\eta_{\rm T}=10^{12}\rm{cm}^2\!/s$ throughout. }
\label{fig:uthrad}
\end{figure}
To obtain a negative radial gradient of $|U_{\theta}|$ we
preserve its general mathematical form (eq. (5) of GRBD), but choose
the parameters of this ansatz as $k=6$ and $n=1.4$
which results in a negative $\partial_r |U_{\theta}|$ within a shallow
surface-near layer of roughly 2\% of the thickness of the convection zone
(see Fig. \ref{fig:uthrad}a). It is worth mentioning
that the stress-free boundary condition
cannot be satisfied with a negative gradient.
With this $U_\theta$ profile the velocities of the magnetic tracers
turn out to be larger than the flow velocity
(red/dot-dashed line in Fig. \ref{fig:uthrad}b). This result
does not come unexpected since the field
lines underneath the surface are pushed northwards by a
flow faster than the surface one. By magnetic tension
the field lines at the surface are then also accelerated.
A snapshot of the magnetic field after two weeks of evolution
is shown in Fig. \ref{fig:brpos}a.
\section{Additional transport mechanisms}
Given that a negative $\partial_r |U_{\theta}|$ does not allow to
reproduce the observations of {\em decelerated} magnetic tracer motion
at the surface, one could be tempted to argue that
the results of the ring diagram analysis could be incorrect.
This conclusion, however, would be premature as our model is by
no means complete. As mentioned in GRBD, studying axisymmetric,
i.e., averaged fields requires the inclusion of the
full mean electromotive force, from which only the turbulent diffusion term
has so far been considered.
What about other constituents?
Under anisotropic conditions, turbulence acts like a mean advective motion,
called {\em turbulent magnetic pumping} and is described by a
vector $\boldsymbol{\gamma}$. In the northern hemisphere,
its $\theta$-component is found to be positive,
i.e., equatorwards whereas its radial component is mainly
negative (downwards), see \cite{KKOS06}.
Being helical, convective turbulence in the rotating Sun exhibits
also an $\alpha$-effect which together with
differential rotation in the uppermost 35Mm of the convection
zone (the so called near-surface shear layer) may well
enable a turbulent dynamo
of the $\alpha\Omega$ type.
As $\partial_r \Omega < 0$ there and $\alpha$ is expected to
be positive in the northern hemisphere,
the dynamo wave should, according to the Parker-Yoshimura rule,
propagate equatorwards, contrary to the MC.
This scenario is difficult to study
as it requires establishing a full nonlinear mean-field dynamo model of
the surface shear layer. Turbulent pumping, in contrast,
can easily be introduced as an additive contribution to the MC.
\begin{figure}
\centering
\vspace{-2mm}
\includegraphics[width=0.7\columnwidth]{bp_pos}
\caption{Magnetic field lines and $r$ component (color coded) after
two weeks propagation, $\eta_{\rm T}=10^{12}\rm{cm}^2\!/s$.
Velocity profile as in Fig.~\ref{fig:uthrad}, red curve.
a) without, b) with pumping, $\gamma_{0 \theta}=10\,\rm{m}/\rm{s}$,
see eq.\eqref{eq:thp}}.
\label{fig:brpos}
\vspace{-5mm}
\end{figure}
We do so employing profiles for $\boldsymbol{\gamma}$
adapted from \cite{GDP08} (hereafter GDP08):
\begin{equation}
\hspace*{-4mm}\begin{alignedat}{3}
\gamma_{\theta} &= &&\gamma_{0 \theta}\left[1+{\rm erf} \!
\left(\frac{r-0.74}{0.035}\right) \right]& \!\!&\left[1-{\rm erf} \!
\left(\frac{r-0.995}{0.05}\right) \right]
\cos\theta \sin^4\theta, \quad \quad \quad \quad \\
\gamma_{r} &= &\,-&\gamma_{0 r}\left[1+{\rm erf}\!
\left(\frac{r-0.7}{0.015}\right) \;\:\right]& \!\!&\left[1-{\rm erf}\!
\left(\frac{r-0.975}{0.1}\right) \right] \!\!
\left[\exp\! \left( \frac{r-0.715}{0.25} \right)^2
\!\!\!\cos\theta +1 \right] \!.
\end{alignedat}\hspace*{-4mm}\label{eq:thp}
\end{equation}
Similar to GDP08, we set $\gamma_{0 r}$ and $\gamma_{0 \theta}$
such that the maximum of $\gamma_{\theta}$ is 2.5 times larger
than that of $\gamma_{r}$.
\footnote{
Note that $\gamma_\theta$ slightly differs from
GDP08 insofar it is here $\ne 0$ at the surface.
In the underlying DNSs a
$\gamma_{\theta}$ vanishing at the surface could be
an artifact due to the boundary conditions.
}
The radial profile of $\gamma_{\theta}$, which is the
more relevant component for this study, is shown with a broken
line in Fig. \ref{fig:uthrad}a.
From direct numerical simulations (DNS) \cite{KKOS06}
have estimated that the maximum
$\gamma_\theta$ should be $\approx2.5$m/s. According
to eq. \eqref{eq:thp} its value
at the surface would then be $1.11$m/s.
Given that the surface amplitude of $|U_\theta|\approx 25$m/s,
this small value cannot significantly modify
the results found in the previous section.
Hence we have increased $\gamma_{0 \theta}$ progressively and found that an
MFT speed profile that roughly coincides with
that of the flow speed is obtained with $\gamma_{0 \theta}=5 \rm{m/s}$
(i.e., $2.2$m/s at the surface or
$\approx10$\% of the flow speed).
For larger values of $\gamma_{0 \theta}$ the MFT speed
is smaller than the flow speed.
In Fig. \ref{fig:uthrad}b we show the result obtained
with the extraordinarily high value $\gamma_{0 \theta}=10$m/s
(see broken lines).
Note that the latitudinal profile of $\gamma_{\theta}$
(eq. \eqref{eq:thp}) peaks at 63$^{\circ}$ while the flow profile
does at 57$^{\circ}$. This is
interesting because the resulting profile of the MFT speed peaks
at a different latitude than the flow,
just as it is obtained in observations (see fig. 10 of
\cite[Ulrich, 2010]{ulrich2010}).
Fig. \ref{fig:brpos} compares the morphology of the magnetic field from
the simulation with pumping (b) to that without (a).
Note that the magnetic field is transported further to the north
in the case without magnetic pumping
\begin{discussion}
Negative gradients $\partial_r |U_\theta|$ at the solar surface,
as suggested by helioseismological ring-diagram analyses,
result in MFT speeds higher than the plasma flow speed.
By enhancing our flux transport model with turbulent magnetic pumping,
we tried to reconcile the helioseismological results with the actually
oberserved {\em lower} MFT speeds.
However, high amplitudes of the pumping velocity not supported by
DNS are needed.
We have to conclude that either: (i) the negative gradient of
$|U_\theta|$, suggested by ring-diagram analysis, if real,
should be smaller than considered here, (ii)
the pumping velocities from DNS are unrealistically low,
which cannot be excluded because the setup in \cite{KKOS06} differs
from the solar conditions, or (iii) the inclusion of the $\alpha$ effect
in the flux transport model, enabling a dynamo process in the Sun's
surface shear layer, is crucial. For the latter,
we have to think of a short-wavelength dynamo wave traveling equatorwards.
Magnetic quenching would cause a corresponding
$\theta$ modulation of $\alpha$, $\boldsymbol{\gamma}$
and $\eta_{\rm T}$ whereas the mean Lorentz force would modulate
$U_\theta$ and $\Omega$.
Here, quenching of $\eta_{\rm T}$ is perhaps most promising as
the value of $\eta_{\rm T}$ ``decides''to what degree the field
is frozen in the fluid.
\end{discussion}
\vspace{2mm}
\noindent
\textit{Acknowledgment}: We
thank A. Mu\~{n}oz-Jaramillo for pointing out the
negative radial gradient of $|U_{\theta}|$ in the ring-diagram analysis and
A. Kosovichev for his valuable comments.
|
\section{Introduction}
\label{sec:intro}
$\Lambda\Lambda$ hypernuclei provide valuable information on the
$\Lambda\Lambda$ interaction and how it fits into our understanding
of the baryon-baryon interaction. Although the existence of
$\Lambda\Lambda$ hypernuclei nearly rules out a stable $H$ dibaryon,
a $\Xi N$ dominated $H$ resonance might affect the systematics of
$\Lambda\Lambda$ binding energies. Only three emulsion events presented
serious candidates for $\Lambda\Lambda$ hypernuclei before 2001:
$_{\Lambda\Lambda}^{~10}$Be \cite{Dan63,DDF89}, $_{\Lambda\Lambda}^{~~6}$He
\cite{Pro66} and $_{\Lambda\Lambda}^{~13}$B \cite{Aok91,DMG91}.
The $\Lambda\Lambda$ binding energies $B_{\Lambda\Lambda}$ deduced from
these events indicated that the ${^{1}S_0}$ interaction $V_{\Lambda\Lambda}$
was strongly attractive, with a $\Lambda\Lambda$ excess binding energy
$\Delta B_{\Lambda\Lambda} \sim 4.5$ MeV, although it had been
realized that the binding energies of $_{\Lambda\Lambda}^{~10}$Be and
$_{\Lambda\Lambda}^{~~6}$He were inconsistent with each other \cite{BUC84}.
Here, the $\Lambda\Lambda$ excess binding energy is defined as
\begin{equation}
\label{eq:delB}
\Delta B_{\Lambda\Lambda} (^{~~A}_{\Lambda \Lambda}Z)
= B_{\Lambda\Lambda} (^{~~A}_{\Lambda \Lambda}Z)
- 2{\bar B}_{\Lambda} (^{(A-1)}_{~~~~\Lambda}Z)\;,
\end{equation}
where
${\bar B}_{\Lambda}$ is the (2$J$+1)-average of
$B_{\Lambda}$ values for the $^{(A-1)}_{~~~~\Lambda}Z$ hypernuclear core
levels. For comparison, $\Delta B_{\Lambda N}({_{\Lambda}^{5}{\rm He}})=
1.73\pm 0.13$ MeV, implying the unnatural ordering $\Delta B_{\Lambda\Lambda}
> \Delta B_{\Lambda N}$. This perception changed in 2001 when a uniquely
assigned $_{\Lambda\Lambda}^{~~6}{\rm He}$ hybrid-emulsion event \cite{Tak01},
with updated values \cite{Nak10}
\begin{equation}
\label{eq:LL6He}
B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}{\rm He})=6.91\pm 0.16~{\rm MeV},
\;\;\;\; \Delta B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}{\rm He})=0.67\pm
0.17~{\rm MeV} \;,
\end{equation}
ruled out the high value of $\Delta B_{\Lambda\Lambda}$ from the dubious
earlier $_{\Lambda\Lambda}^{~~6}{\rm He}$ event \cite{Pro66}, restoring thus
the expected hierarchy $\Delta B_{\Lambda\Lambda} < \Delta B_{\Lambda N}$.
Both capture at rest formation $\Xi^- +{^{12}{\rm C}}\to{_{\Lambda\Lambda}^
{~~6}{\rm He}} + t + \alpha$ and weak decay $_{\Lambda\Lambda}^{~~6}
{\rm He}\to {_{\Lambda}^{5}{\rm He}}+p+\pi^-$, in this so called NAGARA event,
yield consistently with each other the values listed in (\ref{eq:LL6He}).
Neither $_{\Lambda\Lambda}^{~~6}{\rm He}$ nor $_{\Lambda}^{5}{\rm He}$
have excited states that could bias the determination of $B_{\Lambda\Lambda}
(_{\Lambda\Lambda}^{~~6}{\rm He})$.
Accepting the NAGARA event calibration of $V_{\Lambda\Lambda}$, we review
and discuss (i) particle stability for lighter $\Lambda\Lambda$ hypernuclei;
(ii) reinterpretation of the events assigned $_{\Lambda\Lambda}^{~10}{\rm Be}$
and $_{\Lambda\Lambda}^{~13}{\rm B}$; (iii) several alternative assignments
for the $_{\Lambda\Lambda}^{~13}{\rm B}$ event;
and (iv) plausibility of the assignments $_{\Lambda\Lambda}^{~11}{\rm Be}$
or $_{\Lambda\Lambda}^{~12}{\rm Be}$ proposed for the recently reported HIDA
event \cite{Nak10}. In the course of doing so, we compare $B_{\Lambda\Lambda}$
values derived from emulsion events with shell-model predictions \cite{GMi11}
and with selected few-body cluster calculations \cite{BUC84,HKM02,HKY10} where
the latter exist.
\section{Onset of $\Lambda\Lambda$ hypernuclear stability}
\label{sec:onset}
From the very beginning it was recognized that $\Lambda\Lambda$ and
$\Lambda\Lambda N$ were unbound \cite{Dal63b,THe65}; if $\Lambda\Lambda N$
were bound, the existence of a $nn\Lambda$ bound state would follow.
The existence of a $_{\Lambda\Lambda}^{~~4}$H bound state was claimed by
AGS-E906 \cite{Ahn01a}, from correlated weak-decay pions emitted sequentially
by $\Lambda\Lambda$ hypernuclei produced in a $(K^-,K^+)$ reaction on $^9$Be.
However, the $_{\Lambda\Lambda}^{~~4}$H interpretation is controversial
\cite{KFO02,RHu07}. Several post-2001 calculations exist for
$^{~~4}_{\Lambda\Lambda}$H. A Faddeev-Yakubovsky 4-body calculation finds
no bound state \cite{FGa02c}, whereas a stochastic-variational (SV) 4-body
calculation finds it to be bound by as much as 0.4 MeV \cite{NAM03}.
The more comprehensive $s$-shell $\Lambda$- and $\Lambda\Lambda$-hypernuclear
SV calculation of Ref.~\cite{NSA05} finds $^{~~4}_{\Lambda\Lambda}$H to be
particle stable by as little as a few keV, which would be insufficient to
maintain particle stability once $V_{\Lambda\Lambda}$ is renormalized to
reproduce the recently updated (smaller) value of Eq.~(\ref{eq:LL6He}) for
$\Delta B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}{\rm He})$.
\begin{figure*}[htb]
\includegraphics[width=0.7\textwidth]{e6e5.eps}
\caption{Faddeev calculations of
$\Delta B_{\Lambda\Lambda}({_{\Lambda\Lambda}^{~~6}{\rm He}})$ {\it vs.}
$\Delta B_{\Lambda\Lambda}({_{\Lambda\Lambda}^{~~5}{\rm H}},
{_{\Lambda\Lambda}^{~~5}{\rm He}})$ \cite{FGa02b}, see text.}
\label{fig:2}
\end{figure*}
Regardless of whether $^{~~4}_{\Lambda\Lambda}$H is particle-stable or not,
there is a general consensus that the mirror $\Lambda\Lambda$ hypernuclei
$^{~~5}_{\Lambda\Lambda}$H--$^{~~5}_{\Lambda\Lambda}$He are particle-stable,
with $\Delta B_{\Lambda\Lambda}\sim 0.5-1$ MeV \cite{FGa02b,FGS03}, or larger
owing to the $\Lambda\Lambda-\Xi N$ coupling which is particularly effective
here \cite{NSA05,MSA03,LYa04}. In addition, substantial charge symmetry
breaking effects are expected in these systems, resulting in a higher
binding energy of $^{~~5}_{\Lambda\Lambda}$He by up to 0.5 MeV with respect
to $^{~~5}_{\Lambda\Lambda}$H \cite{LYa04,YRi08}. Figure~\ref{fig:2}
demonstrates how $\Delta B_{\Lambda\Lambda}$ values for the $A=5,6$ systems,
calculated over a broad range of $V_{\Lambda\Lambda}$ strengths, are nearly
linearly correlated with only a small offset. Thus, the stability of
$^{~~6}_{\Lambda\Lambda}$He ensures stability for $^{~~5}_{\Lambda\Lambda}$H.
\section{Ingredients of hypernuclear shell model}
\label{sec:SM}
Shell-model predictions for $\Lambda\Lambda$ hypernuclei have been
given recently \cite{GMi11} using Eq.~(\ref{eq:delB}) in which
$\Delta B_{\Lambda\Lambda}(^{~~A}_{\Lambda \Lambda}Z)$ is replaced by
a constant $V_{\Lambda\Lambda}$ matrix element, identified with
$\Delta B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}{\rm He})$ of
Eq.~(\ref{eq:LL6He}).{\footnote{A straightforward modification for
$_{\Lambda\Lambda}^{~10}{\rm Be}_{\rm g.s.}(0^+)$, with a nuclear core
$^8$Be unstable to $\alpha$ emission, is discussed in Ref.~\cite{GMi11}.}}
Calculations of $B_{\Lambda\Lambda}(^{~~A}_{\Lambda\Lambda}Z)$ require then
the knowledge of ${\bar B}_{\Lambda}(^{(A-1)}_{~~~~\Lambda}Z)$ involving
single-$\Lambda$ hypernuclear ground-state (g.s.) binding energies plus
g.s. doublet splittings $\Delta E_{\rm g.s.}$ for $J_{\rm core}\neq 0$.
Table~\ref{tab:1} lists $\Delta E_{\rm g.s.}$ values relevant for the
calculations reviewed here, exhibiting remarkable agreement between
theory and experiment.
\begin{table}[htb]
\caption{Doublet splittings $\Delta E^{\rm th}$ and $\Delta E^{\rm exp}$
(in keV) from Refs.~\cite{Mil10,Tam05}, where $\Delta E^{\rm th}_{\rm alt}$
uses ESC04a--inspired $\Lambda-\Sigma$ coupling. Note the sensitivity
of $\Delta E^{\rm th}({_{~\Lambda}^{10}{\rm B}_{\rm g.s.}})$ to the model
used for $\Lambda-\Sigma$ mixing. The $_{\Lambda}^{9}{\rm Be}^{\ast}$
and $_{~\Lambda}^{13}{\rm C}^{\ast}$ excited doublets are discussed in
Sect.~\ref{sec:consistency}.}
\label{tab:1}
\begin{tabular}{lccccc}
\hline\noalign{\smallskip}
& $J_{\rm up}^{\pi}$ & $J_{\rm low}^{\pi}$ & $\Delta E^{\rm th}$ &
$\Delta E^{\rm th}_{\rm alt}$ & $\Delta E^{\rm exp}$ \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$_{\Lambda}^{9}{\rm Be}^{\ast}$ & $3/2^+$ & $5/2^+$ & 44 & 49 & 43$\pm$5 \\
\noalign{\smallskip}
$_{~\Lambda}^{10}{\rm B}_{\rm g.s.}$ & $2^-$ & $1^-$ & 120 & 34 &
$\leq 100$ \\
\noalign{\smallskip}
$_{~\Lambda}^{11}{\rm B}_{\rm g.s.}$ & $7/2^+$ & $5/2^+$ & 267 & 243 &
262.9$\pm$0.2 \\
\noalign{\smallskip}
$_{~\Lambda}^{12}{\rm C}_{\rm g.s.}$ & $2^-$ & $1^-$ & 153 & 167 &
161.4$\pm$0.7 \\
\noalign{\smallskip}
$_{~\Lambda}^{13}{\rm C}^{\ast}$ & $5/2^+$ & $3/2^+$ & 31 & 47 & -- \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
\section{Interpretation of $_{\Lambda\Lambda}^{~10}{\rm Be}$
and $_{\Lambda\Lambda}^{~13}{\rm B}$ emulsion events
}
\label{sec:consistency}
The $B_{\Lambda\Lambda}$ values of both $_{\Lambda\Lambda}^{~10}{\rm Be}$
($17.5\pm 0.4$ MeV) \cite{DDF89} and $_{\Lambda\Lambda}^{~13}{\rm B}$
($28.2\pm 0.7$ MeV) \cite{Aok09} were extracted assuming that their $\pi^-$
weak decay proceeds to the g.s. of the respective daughter
$\Lambda$ hypernuclei. This led to $\Delta B_{\Lambda\Lambda}\sim 4-5$ MeV,
substantially higher than for $_{\Lambda\Lambda}^{~~6}{\rm He}$ (NAGARA).
However, as realized by Danysz {\it et al.} \cite{Dan63}, the decay could
proceed to excited states of the daughter $\Lambda$ hypernucleus which
deexcites then rapidly to the g.s. emitting unobserved $\gamma$ radiation.
This reduces the apparent $B_{\Lambda\Lambda}$ and $\Delta B_{\Lambda\Lambda}$
values by the $\Lambda$ hypernuclear excitation energy involved in the $\pi^-$
weak decay. Consistency with $_{\Lambda\Lambda}^{~~6}{\rm He}$ is restored
upon accepting the following weak decays:
\begin{equation}
_{\Lambda\Lambda}^{~10}{\rm Be} \rightarrow {_{\Lambda}^{9}{\rm Be}}^{\ast}
(3/2^+,5/2^+;3.04~{\rm MeV}) + p + \pi^- ,
\label{eq:exc9}
\end{equation}
\begin{equation}
_{\Lambda\Lambda}^{~13}{\rm B} \rightarrow {_{~\Lambda}^{13}{\rm C}}^{\ast}
(3/2^+,5/2^+;4.9~{\rm MeV}) + \pi^-,
\label{eq:exc13}
\end{equation}
with rates comparable to those for decays to $_{\Lambda}^{9}{\rm Be}_{\rm g.s.}
(1/2^+)$ and $_{~\Lambda}^{13}{\rm C}_{\rm g.s.}(1/2^+)$, respectively.
The doublet splittings of $_{\Lambda}^{9}{\rm Be}^{\ast}$ and
$_{~\Lambda}^{13}{\rm C}^{\ast}$ are listed in Table~\ref{tab:1}.
$_{\Lambda\Lambda}^{~10}{\rm Be}$ also fits the Demachi-Yanagi event
observed in KEK-E373 \cite{Ahn01b}, with $B_{\Lambda\Lambda}=11.90\pm 0.13$
MeV \cite{Nak10} determined from the assumed formation reaction kinematics.
The approximately 6 MeV difference between this and the
Danysz {\it et al.} \cite{Dan63,DDF89} value for
$B_{\Lambda\Lambda}({_{\Lambda\Lambda}^{~10}{\rm Be}})$ is reconciled by
assuming that the Demachi-Yanagi event corresponds to formation of the
first excited state $_{\Lambda\Lambda}^{~10}{\rm Be}^{\ast}$,
\begin{equation}
\Xi^- + {^{12}{\rm C}} \rightarrow
{_{\Lambda\Lambda}^{~10}{\rm Be}^{\ast}(2^+;\approx 3~{\rm MeV})} + t,
\label{eq:exc10}
\end{equation}
which decays to $_{\Lambda\Lambda}^{~10}{\rm Be}_{\rm g.s.}$ by emitting
unseen $\gamma$ ray, the energy of which has to be added to the apparent
$B_{\Lambda\Lambda}$ value deduced by assuming a g.s. formation.
It is not clear why the formation of $_{\Lambda\Lambda}^{~10}{\rm Be}^{\ast}$
should be comparable or enhanced with respect to that of
$_{\Lambda\Lambda}^{~10}{\rm Be}_{\rm g.s.}$.
The $B_{\Lambda\Lambda}^{\rm exp}$ values corresponding to
Eqs.~(\ref{eq:exc9})--(\ref{eq:exc10}) are listed in Table~\ref{tab:2}
together with predictions made in cluster model (CM) and shell model (SM)
calculations, all of which use $\Lambda\Lambda$ interactions normalized to
$B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}{\rm He})=6.91\pm 0.16$ MeV
\cite{Nak10}. For $_{\Lambda\Lambda}^{~13}{\rm B}$, assuming charge symmetry,
the $_{~\Lambda}^{12}{\rm B}_{\rm g.s.}$ doublet splitting input was
identified with that of $_{~\Lambda}^{12}{\rm C}_{\rm g.s.}$ from
Table~\ref{tab:1}. It is seen that both CM and SM calculations reproduce the
reinterpreted $B_{\Lambda\Lambda}$ values of $_{\Lambda\Lambda}^{~10}{\rm Be}$
and $_{\Lambda\Lambda}^{~13}{\rm B}$. The SM agrees well with the
Hiyama {\it et al.} CM calculation \cite{HKM02,HKY10}, and the SM
calculation has no match for $_{\Lambda\Lambda}^{~13}{\rm B}$.
\begin{table}[htb]
\caption{Reinterpreted $B_{\Lambda\Lambda}^{\rm exp}$ values (in MeV) and
predictions based on the NAGARA event for $_{\Lambda\Lambda}^{~~6}{\rm He}$.
The error on $B_{\Lambda\Lambda}^{\rm exp}({_{\Lambda\Lambda}^{~~6}{\rm He}})$
is incorporated into the predicted values.}
\label{tab:2}
\begin{tabular}{cccccc}
\hline\noalign{\smallskip}
& \multicolumn{2}{c}{$B_{\Lambda\Lambda}^{\rm exp}$} &
\multicolumn{2}{c}{$B_{\Lambda\Lambda}^{\rm CM}$} &
\multicolumn{1}{c}{$B_{\Lambda\Lambda}^{\rm SM}$} \\
& Eqs.~(\ref{eq:exc9},\ref{eq:exc13})& Eq.~(\ref{eq:exc10}) &
\cite{BUC84} & \cite{HKY10} & \cite{GMi11} \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$_{\Lambda\Lambda}^{~10}{\rm Be}$ & $14.5\pm 0.4$ & $14.94\pm 0.13$ &
$14.35\pm 0.19$ & $14.74\pm 0.19$ & $14.97\pm 0.22$ \\
\noalign{\smallskip}
$_{\Lambda\Lambda}^{~13}{\rm B}$ & $23.3\pm 0.7$ & &--&--&
$23.21\pm 0.21$ \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
\section{Alternative interpretations of the $_{\Lambda\Lambda}^{~13}{\rm B}$
event}
\label{sec:L13B}
The emulsion event assigned to $_{\Lambda\Lambda}^{~13}{\rm B}$
\cite{Aok91,DMG91} has been carefully scrutinized by the KEK-E176
Collaboration \cite{Aok09}. Several alternative assignments were pointed out,
two of which that do not require $\Lambda$ hypernuclear excitation in the
$\pi^-$ weak decay of the $\Lambda\Lambda$ hypernuclear g.s. are listed in
Table~\ref{tab:3}. Comparison with model calculations suggests that such
reassignments cannot be ruled out, although a $_{\Lambda\Lambda}^{~13}{\rm B}$
assignment shows a higher degree of consistency between the
$B_{\Lambda\Lambda}$ values derived from formation and from decay.
In particular, the accepted formation reaction $\Xi^- + {^{14}{\rm N}} \to
{_{\Lambda\Lambda}^{~13}{\rm B}} + p + n$ was shown to occur naturally in
$\Xi^-$ capture at rest in light nuclei emulsion \cite{DMG91}.
\begin{table}[htb]
\caption{Reassignments of the $_{\Lambda\Lambda}^{~13}{\rm B}$
KEK-E176 event. $B_{\Lambda\Lambda}$ values are in MeV.}
\label{tab:3}
\begin{tabular}{cccc}
\hline\noalign{\smallskip}
& $B_{\Lambda\Lambda}^{\rm exp}$ \cite{Aok09} & $B_{\Lambda\Lambda}^{\rm CM}$
\cite{HKY10} & $B_{\Lambda\Lambda}^{\rm SM}$ \cite{GMi11} \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$_{\Lambda\Lambda}^{~11}{\rm Be}$ & $17.53\pm 0.71$ & $18.23\pm 0.19$ &
$18.40\pm 0.28$ \\
\noalign{\smallskip}
$_{\Lambda\Lambda}^{~12}{\rm B}$ & $20.60\pm 0.74$ & -- & $20.85\pm 0.20$ \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
\section{Interpretation of the KEK-E373 HIDA event}
\label{sec:HIDA}
\begin{table}[htb]
\caption{Assignments suggested for the KEK-E373 HIDA event.
$B_{\Lambda\Lambda}$ values are in MeV.}
\label{tab:4}
\begin{tabular}{cccc}
\hline\noalign{\smallskip}
& $B_{\Lambda\Lambda}^{\rm exp}$ \cite{Nak10} & $B_{\Lambda\Lambda}^{\rm CM}$
\cite{HKY10} & $B_{\Lambda\Lambda}^{\rm SM}$ \cite{GMi11} \\
\noalign{\smallskip}\hline\noalign{\smallskip}
$_{\Lambda\Lambda}^{~11}{\rm Be}$ & $20.83\pm 1.27$ & $18.23\pm 0.19$ &
$18.40\pm 0.28$ \\
\noalign{\smallskip}
$_{\Lambda\Lambda}^{~12}{\rm Be}$ & $22.48\pm 1.21$ & -- & $20.72\pm 0.20$ \\
\noalign{\smallskip}\hline
\end{tabular}
\end{table}
The KEK-E373 Collaboration has recently presented evidence from the HIDA event
for another $\Lambda\Lambda$ hypernucleus, tentatively assigned to either
$_{\Lambda\Lambda}^{~11}{\rm Be}$ or to $_{\Lambda\Lambda}^{~12}{\rm Be}$
\cite{Nak10}. The associated $B_{\Lambda\Lambda}^{\rm exp}$ values, together
with model predictions, are listed in Table~\ref{tab:4}. We note that since no
experimental data exist on $_{\Lambda}^{11}{\rm Be}$, the required input for
evaluating $B_{\Lambda\Lambda}^{\rm SM}({_{\Lambda\Lambda}^{~12}{\rm Be}})$
was derived within the SM approach \cite{GMi11}. It is clear from the table
that neither of the proposed assignments is favorable, although the relatively
large experimental uncertainties do not completely rule out either of these.
\section{Conclusion}
\label{sec:concl}
It was shown how the three acceptable $\Lambda\Lambda$ emulsion events,
corresponding to $_{\Lambda\Lambda}^{~~6}{\rm He}$,
$_{\Lambda\Lambda}^{~10}{\rm Be}$ and $_{\Lambda\Lambda}^{~13}{\rm B}$,
can be made consistent with each other, in good agreement with CM and with
SM calculations of $B_{\Lambda\Lambda}$. Other possible assignments for
the KEK-E176 $_{\Lambda\Lambda}^{~13}{\rm B}$ event were discussed, and
the assignments proposed for the recently reported HIDA event were found
unfavorable. It was pointed out that simple shell-model estimates, making
use of $\Lambda$-hypernuclear spectroscopic data and analysis, are sufficient
for discussing the world data of $\Lambda\Lambda$ hypernuclear events.
A relatively weak $\Lambda\Lambda$ interaction, with $(1s_{\Lambda})^2$
matrix element of magnitude $\Delta B_{\Lambda\Lambda}(_{\Lambda\Lambda}^{~~6}
{\rm He})=0.67\pm 0.17$ MeV, describes well the data in the observationally
accessible mass range $6 \leq A \leq 13$. Comparably weak $\Lambda\Lambda$
interactions are obtained also in recent theoretical models, in Nijmegen
extended soft-core (ESC) models \cite{YRi08,RNY10} and in lowest order
$\chi$EFT \cite{PHM07}. Less well determined is the $\Lambda\Lambda$ coupling
to the slightly higher $\Xi N$ channel, with appreciable model dependence in
ESC models \cite{YRi08,RNY10}. The observation of $A=5$ $\Lambda\Lambda$
hypernuclei would add valuable new information on this issue.
\begin{acknowledgements}
Useful discussions with Emiko Hiyama on hypernuclear cluster-model
calculations are gratefully acknowledged.
\end{acknowledgements}
|
\section{Introduction}
\vspace{-10pt}
In 1981, Guth~\cite{Guth:1980zm} introduced an inflationary phase (old inflation) to explain the horizon, flatness, and monopole problems.
The Universe begins hot, cools, and becomes
stuck in a false vacuum and begins to inflate.
Eventually,
the Universe transitions to the true vacuum and reheats.
Unfortunately, Guth's model failed due to the Swiss cheese problem or lack of a graceful exit.
The tunneling rate to the true vacuum
must be small to generate a sufficient amount
of inflation, but then inflation never ends.
New inflation~\cite{Linde:1981mu,Albrecht:1982wi}
sidestepped the Swiss cheese problem by introducing a slowly rolling scalar field to drive inflation.
Slow roll not only solves the standard cosmological problems but can also generate adiabatic density perturbations
consistent with the CMB.
Regardless, slow roll has two serious generic drawbacks beyond the fine-tuning problem~\cite{Adams:1990pn};
first, the high scale of inflation (which
prevents testing inflation in the lab) leads to problems from trans-Planckian physics to overclosure from moduli and gravitinos; second (a much worst
problem),
eternal inflation
(which leads to the measure problem)
undermines the predictive power of inflation. Hence, one is interested in alternatives to slow roll
such as cyclic models and ekpyrotic Universes~\cite{Khoury:2001wf}, but these models must contend with a singular bounce,
which introduces a different can of worms.
\begin{figure}[t]
\includegraphics[width=0.35\textwidth]{Apexinfcol.pdf}
\centering
\caption{{\bf top--} The Universe is stuck in a false vacuum and supercools. {\bf bottom--} At T$_c$, a Yukawa term
generates a tadpole term; the barrier goes away. The Universe then rapidly transitions
to the true vacuum.}
\vspace{-15pt}
\label{apexinf}
\end{figure}
Instead, we do away with any scalar dynamics to control inflation.
A thermal bath (present during old inflation) regulates inflation.
We introduce a new
thermal and technically natural mechanism to generate a graceful exit.
We have dubbed the model SuperCool (SC) inflation since the Universe supercools during inflation and then rapidly transitions
to the true vacuum due to a small perturbation,
in much the same that a supercooled liquid almost instantaneously freezes if slightly disturbed. The model in spirit is similar to thermal
inflation~\cite{Lyth:1995ka} except unlike thermal inflation
our model successfully solves the cosmological problems and generates adiabatic density perturbations.
SuperCool Inflation (SCI) works at the TeV scale and below and avoids eternal inflation.
For simplicity, SuperCool (SC) field is a complex scalar with a Coleman-Weinberg potential
\begin{equation}
\label{fp}
\begin{aligned}
&V(\phi)=(1/8)\,\text{T}^2\,\phi^2+\frac{3g_X^4}{32 \pi^2}|\phi|^4\bigg(\ln\bigg(\frac{|\phi|}{\langle|\phi|\rangle}\bigg)-\frac{1}{4}\bigg)\\
&+\sum y_i \phi\ast (q_R \bar q_L)_i+ \text{h.c.} +\Lambda^4+\text{ non-renormalizable}
\end{aligned}
\end{equation}
charged under a
U(1)$_X$ gauge group with a charge $g_x$, where $\Lambda^4$ is the standard cosmological tuning, which tunes the cosmological
constant of the true vacuum to zero. Finite temperature effects generate an effective mass term (T$^2\phi^2$).
The SC field couples to a set of QCD like fermions $q_R \bar q_L$ with a Yukawa coupling $\lambda_i$.
The SC sector is a simplified version of the standard model with the SC inflaton mapped to the Higgs boson. The
SU(2) weak force accompanied by leptons has been dropped.
The non-renormalizable terms will play an important role in avoiding eternal inflation.
We will discuss Eq.~\ref{fp} in detail in the text but first describe the qualitative behavior of Eq.~\ref{fp}.
In the beginning, the Universe is hot and dense, becomes stuck in a false vacuum (Fig.\ \ref{apexinf}{--top}), and
inflates solving the horizon and flatness problems.
The temperature
of the Universe falls exponentially. In section \ref{stabilized},
we show that finite temperature effects stabilize the potential against tunneling during inflation.
In section \ref{witten}, we introduce a technically natural way to end inflation.
When the Universe reaches the critical temperature T$_c$, a non-Abelian gauge group freezes,
which triggers the end of inflation.
A set of QCD like fermions charged under the
non-Abelian gauge group get a vev $\langle \lambda \bar\lambda\rangle\simeq$ T$_c^3$. The fermions
(which have a Yukawa coupling to the SC field)
generate a tadpole term for the inflaton potential. The new term removes the
barrier trapping the scalar field in the false vacuum (Fig.\ \ref{apexinf}{--bottom}).
We remind the reader that as the temperature of the Universe drops the barrier trapping the false becomes smaller;
the barrier scales with T.
The field then quickly rolls down to the true vacuum and reheats the Universe.
The mechanism is technically natural; the end of inflation is determined by the logarithmic
running of a non-Abelian gauge group. In the same way, the QCD scale is technically natural compared to the Planck scale,
which is a difference of $10^{20}$ orders of magnitude and many orders of magnitude more than
the temperature difference between when SC inflation begins and ends.
In the case of radiative Electro-Weak (EW)
symmetry breaking, Weinberg and Guth~\cite{Guth:1981uk} noted that the Universe would become trapped in
the symmetric false vacuum of the Higgs potential to arbitrarily low temperatures.
Witten~\cite{Witten:1980ez} showed that when the QCD vacuum freezes,
the Yukawa terms generate a tadpole term for the Higgs potential. The new term
destabilizes the false vacuum and the Universe transitions to the true vacuum. We take Witten's
observation and introduce it as a way to generate a graceful exit for old inflation.
After the field goes to the true vacuum in section \ref{reheating},
we show that the Universe rapidly reheats. We discuss the different ways that the inflaton can couple
to the standard model. As an instructive example, we consider kinetic mixing between hypercharge U(1)$_Y$
and U(1)$_{X}$ which generates a $Z^\prime$. The SC inflaton then decays into a pair of Z bosons.
A high scale of inflation (order GUT) can be problematic.
The GUT scale generically results from the
dual necessity to generate a sufficient number of efolds and
to generate the correct spectrum of density perturbations~\cite{Adams:1990pn}.
If the height of the potential is on the order of the GUT scale, then the width frequently needs to be larger
than the Planck scale, at which point we must deal with trans-Planckian physics. At a practical level,
non-renormalizable terms suppressed by $m_{pl}$ can become large
and dangerous~\cite{1997PhRvL..78.1861L}. High scale models also run into overclosure problems
from moduli and gravitinos.
From a collider perspective, a high scale of inflation is disappointing without a GUT scale collider.
In contrast in section \ref{scale}, we show that SC inflation occurs at the TeV scale and down avoiding the complications of high scale inflation.
Previous authors have proposed inflation at the TeV scale. In fact a model introduced by Turner \& Knox~\cite{Knox:1992iy}
similarly uses a Coleman Weinberg (CW) potential (We use a CW potential for the SC field),
but all of the models are rolling field models. TeV scale rolling models
often have difficulties from fine tuning problems~\cite{Lyth:1999ty}
or require unusual initial conditions~\cite{German:2001tz}.
SC inflation itself does not suffer from fine-tuning, but our proposed mechanism to
generate perturbations (which is similar to a curvaton) suffers from a potential
fine-tuning which we discuss in section \ref{perturbations}.
Clearly, we see the interest in connecting TeV scale inflation with EW symmetry breaking
or more generally with beyond the Standard Model (SM) building such as hidden valley models~\cite{Strassler:2006im} etc..
We have not attempted to directly connect SC inflation and EW symmetry breaking, but the path is clear.
One can write down a gauge invariant dimension 4 renormalizable scalar coupling between the SM Higgs and SC field.
When the SC field gets a vev, it can generate a negative mass term which induces spontaneous EW symmetry breaking.
We have left detailed model building to future work. In that spirit, we have written the paper from an effective field theory perspective
without reference to any particular fundamental theory which might relate to SUSY, GUTs, etc..
Next, we turn to eternal inflation in section \ref{eternal}, which is endemic of the slow roll paradigm and troublesome.
Once one has gone to the trouble of constructing a slow
roll potential, it has been recently shown that the potential must eternally inflate~\cite{2011PhRvD..84b3511K}.
In an eternally inflating Universe (a Multiverse for short),
most of the Universe is stuck inflating.
Regardless, small pockets of the Multiverse will stop inflating and
could be like our visible Universe or completely different.
Critically, we need to measure the relative probability of different pocket Universes to make predictions.
After 25 years of extraordinary effort, no measure predicts a Universe which looks at all like our own.
For instance,
the geometric or light cone measure (introduced by Bousso~\cite{Bousso:2006ev}) predicts
that time itself will end in the next 5 Billion years~\cite{Bousso:2010yn}.
We list a few references
~\cite{Nomura:2011dt,Harlow:2011az,Garriga:2008ks,
Bousso:2010im,Linde:2008xf,DeSimone:2008if,Vilenkin:2011yx}
of the extensive literature on the measure problem.
In section \ref{eternal}, we show that SCI avoids eternal inflation.
In general, any scalar field has two regions in which eternal inflation can crop up:
at a hilltop of a potential as with the original new inflation
model or at large field values as with chaotic inflation.
The SC inflaton potential has neither of these troublesome points.
A non-minimal coupling to gravity or non-renormalizable terms
prevent the large field value case. At the hilltop points in SCI,
the inflaton does not slow roll, in which case there is no
eternal inflation as shown in the Appendix. In fact, one would need
to introduce a large tuning to make the hilltop sufficiently flat to have
slow roll in the first place.
We introduce in the last major section \ref{perturbations} a novel way of generating perturbations.
The thermal background which controls SCI suppresses
normal density perturbations from a scalar field. We will require a new way to generate perturbations.
Isocurvature (entropy) perturbations can generate real adiabatic density perturbations. Mollerach showed that if
a matter component (which has an isocurvature perturbation) decays into radiation, then the matter component's
isocurvature perturbation will become a real adiabatic density perturbation~\cite{Mollerach:1989hu}.
The curvaton model~\cite{Lyth:2001nq,Lyth:2002my} has implemented Mollerach's original idea.
We similarly take advantage of Mollerach's mechanism except our model works at much lower inflationary scales compared to the curvaton.
In Section \ref{aulosmechanism}, we introduce the aulos
field (a pseudo Nambu-Goldstone boson).
Aulos is the Greek root for a flute or reed instrument.
The field generates real density perturbations, which seed the formation of all known structure in the Universe.
The ancient Greeks described the motion of celestial objects as the music of the spheres.
Aulos in a similar spirit refers to a source of cosmic harmony.
After spontaneously breaking the U(1) aulos symmetry in section \ref{aulosmass}, we generate a small technically natural mass
by explicitly breaking the U(1) symmetry.
Then in \ref{aulsoevolution}, we discuss the evolution of the aulos field.
During inflation, the aulos field is Hubble damped, but De Sitter fluctuations
induce spatial variation of the misalignment angle of the aulos field. During inflation the mass of the aulos field ($m_a$) is
smaller than the Hubble parameter. The aulos decay constant ($f_a$) is within $\mathcal{O}(10^4)$ of
the Hubble parameters. At the end of inflation, both $m_a$ and $f_a$ grow and become substantially larger than H,
which transfers energy from the inflaton field into the aulos field.
The aulos field begins to
oscillate, which generates a cold condensate of the aulions. Now, the spatial variation of the misalignment
angle of the aulos field induces an isocurvature perturbation. The aulos field then
decays into radiation and generates real density perturbations.
Next, we show in section \ref{signatures} that
the aulos mechanism can produce the perturbations seen in the CMB
and potentially generate some novel features.
The aulos field is very similar to the curvaton scenario except the aulos mechanism works with inflationary scales much smaller than $10^9$ GeV.
Hence, many of the curvaton features
apply to the aulos scenario. First, the aulos mechanism
generates the scale of perturbations seen in the CMB and the spectral tilt
of the power spectrum. Second, the very low scale of inflation
suppresses primordial gravity B--modes.
The aulos field can also potentially generate levels of non-Gaussianities
and isocurvature perturbations, which are observable with Planck and other future missions.
In section \ref{collider}, we then connect the cosmological parameters to the
parameters of the aulos field which could be measured in a lab.
We would like to point out that the aulos mechanism potentially has applications beyond providing perturbations
during inflation. We define an aulos field as any field
which has a small decay constant ( spontaneous symmetry breaking scale ) and/or mass
during inflation. At the end of inflation, the field has a large decay constant and/or mass.
We will discuss some potential applications in the conclusions from variations in dark energy to spatial variation in $\alpha$~\cite{doug1}.
It is important to emphasize despite the many different issues discussed in the paper that the underlying idea is simple. If we can
free ourselves of requiring the inflaton to generate perturbations, we can do away with slow roll altogether and the
many complications which result from slow roll from technical concerns about fine-tuning to more prosaic concerns about the measure problem.
We can also think seriously about low scale inflation.
\section{SuperCool Inflation}\label{SCinflation}
\begin{figure}[t]
\includegraphics[width=0.35\textwidth]{p1.pdf}
\centering
\caption{Coleman Weinberg potential for the SC field $\phi$ (See Eq.~\ref{CW}), at zero temperature. }
\vspace{-15pt}
\label{CWfig}
\end{figure}
As our starting point, the SuperCool (SC) field $\phi$ has a Coleman Weinberg (CW) potential
(See Fig.~\ref{CW}) at zero temperature. Furthermore, SC field is a complex scalar, which is charged under a
U(1)$_X$ Abelian group. We have enforced classical scale invariance on the potential ( known as the ``no bare mass"
condition)\footnote{The naturalness of Eq.~\ref{CWcondition} is open to question, but Gildener and Weinberg have argued that
the ``no bare mass" condition is natural~\cite{Gildener:1976ih}. In addition, CW potential have been used
extensively for EW symmetry breaking and slow roll inflation.
At the minimum, we take classical scale invariance (``no bare mass" condition) as an interesting hypothesis
and continue.} such that
\begin{equation}
\label{CWcondition}
\frac{d^2\text{V}(\phi)}{d\phi^2}\bigg\vert_{\phi=0}=0.
\end{equation}
Classically, the SC field is simply of the form $\lambda\phi^4$ and is scale invariant. Quantum corrections modify the classical potential
breaking the scale invariance. Then, dimensional transmutation generates a nonzero vacuum expectation value $\langle\phi\rangle$ i.e.\ vev.
At one loop, the potential~\cite{Coleman:1973jx} is
\begin{equation}
\label{CW}
\text{V}(\phi)=\frac{3g_X^4}{32 \pi^2}|\phi|^4\bigg(\ln\bigg(\frac{|\phi|}{\langle|\phi|\rangle}\bigg)-\frac{1}{4}\bigg)+ \Lambda^4
\end{equation}
where $g_X=0.4$ is the charge of the scalar field. $\Lambda$ is the usual cosmological tuning, which will generate the
vacuum energy during inflation V$(0)=\Lambda^4$ and sets the true vacuum to zero V$(\langle \phi\rangle)=0$.
We will take $\Lambda\simeq1$ TeV and $\langle|\phi|\rangle\simeq10$ TeV to be concrete.
Clearly, the vev $\langle\phi\rangle$ and the scale of inflation are connected for a given $g_X$.
The scale of inflation i.e.\ $\Lambda$ can be larger or smaller than 1 TeV;
we will comment later. So far, we have neglected temperature effects.
\begin{figure*}
\includegraphics[width=.7\textwidth]{vert2.pdf}
\centering
\caption{The effective potential at different temperatures:
{\bf a}--At high temperatures, only one vacuum state exist.
{\bf b}--Eventually, other vacuum states appear.
{\bf c}--At the transition temperature, the Universe will be in the symmetric minima.
{\bf d}--As the temperature drops, the symmetric vacuum remains metastable.}
\label{tempfig}
\end{figure*}
\subsection{Stabilized Potential}\label{stabilized}
At a nonzero temperature, finite temperature effects stabilize the false vacuum of the SC potential.
Near the origin of $\phi$, we can approximate
the finite temperature part of the effective potential with
\begin{equation}
V(\text{T}, \phi)=\mathcal{O}(1)\text{T}^4+(1/8)\,\text{T}^2\,\phi^2+\mathcal{O}(\phi^4)
\end{equation}
(\cite{Witten:1980ez,Dolan:1973qd,Weinberg:1974hy}). At very high temperatures (Fig.\ref{tempfig}-a), the symmetric minimum is the only
vacuum state and the Universe sits at $\phi=0$. As
the temperature of the Universe drops (Fig.\ref{tempfig}-b), the true vacuum appears near $\phi=\langle\phi\rangle$.
The potential evolves only gradually from high temperatures. As the Universe continues
to cool reaching the transition temperature (Fig.\ref{tempfig}-c),
the Universe will still be in the symmetric minima~\cite{Guth:1980zk,Witten:1980ez}.
At temperatures below the transition temperature (Fig.\ref{tempfig}-d), thermal tunneling can allow the field
to transition to the true vacuum, but the rate is exponentially small.
We will follow an argument first given by Witten~\cite{Witten:1980ez}.
The tunneling rate $\Gamma$ will be dominated by the O(3) thermal instanton S$_3$~\cite{Linde:1981zj}.
We have also considered an O(4) instanton~\cite{Linde:1981zj}
and a Hawking-Moss instanton~\cite{Hawking:1981fz}, which are subdominant.
\begin{equation}
\label{tunnelingrate}
\Gamma\sim \text{T}^4 \bigg(\frac{\text{S}_3(\text{T},\phi)}{2\pi\, \text{T}}\bigg)^{3/2} \exp[{-\frac{\text{S}_3}{\text{T}}}]
\end{equation}
where T is the temperature of the Universe. Near the origin and the barrier (the part of the potential relevant
for tunneling), we can transform Eq.~\ref{CW} with
a trick invented by Witten~\cite{Witten:1980ez}
\footnote{Witten shows that near the origin and the barrier, which is the relevant part of
the potential for tunneling that
$\ln(\phi/\langle\phi\rangle)-1/4 \longrightarrow-\ln(\text{M}/\text{T})+\mathcal{O}(1)$.}
\begin{equation}
\label{approx}
\text{V}(\text{T},\phi)=\frac{\text{T}^2}{8}|\phi|^2-\frac{3 g_X^4}{32 \pi^4}|\phi|^4\bigg(\ln\bigg(\frac{\text{M}_{X}}{\text{T}}\bigg)
+\mathcal{O}(1)\bigg)
\end{equation}
where M$_{X}=g_X\langle\phi\rangle$ is the mass of the U(1)$_X$ gauge boson. With Eq.~\ref{CW} now in the
form of Eq.~\ref{approx},
we have an exact solution of the O(3) instanton.
\begin{equation}
\label{S_3}
\text{S}_3=\frac{4 \pi^2}{3} \frac{\text{T}}{g_X^3\ln(\text{M}_X/\text{T})} 19
\end{equation}
where the factor of 19 is a geometric factor.
In SCI, the Universe is stable against tunneling to the true vacuum. At most, only some
small part of the Universe can transition to the true vacuum.
Guth and Weinberg~\cite{Guth:1982pn} have shown that the tunneling rate (per unit time per unit volume) $\Gamma$
compared to the Hubble 4-Volume
\begin{equation}
\label{trate}
\beta=\Gamma/\text{H}^4\gsim9/4\pi=\beta_c,
\end{equation}
must be larger than $\beta_c$ for the Universe to transition from the false to the true vacuum.
We show that $\beta$ for the SC field is much smaller than $\beta_c$ during inflation. Hence, the Universe is stuck in the false vacuum state.
Rapid tunneling to the true
vacuum can only occur when S$_3\rightarrow0$. Upon inspection, only as $g_X$ becomes large or
T goes to zero does S$_3\rightarrow0$, but we show that in either case tunneling will still be extremely small. Hence, a new mechanism will
be needed to generate a graceful exit.
First, $g_X$ never becomes large and stays perturbatively small as the temperature drops
from the TeV scale down to a fraction of an electron volt (T$\leq 10^{-2}$eV). At which point in the next section, we show
that the Witten mechanism can generate a graceful exit. More generally if the SC field was charged under a non-Abelian gauge group,
we would need to be more careful (See \cite{PhysRevD.24.1699}).
Second as the temperature drops, the logarithm $\ln(\text{T}/\text{M}_{X})$ in Eq.~\ref{S_3} blows up. Regardless $\beta$ (for the inflaton potential)
is much less than
$\beta_c$ during inflation, since
the pre-factor in Eq.~\ref{tunnelingrate}
(which scales like T$^4$) goes to zero sufficiently fast to counteract the Log factor in the exponential.
Furthermore, $\beta$ is sufficiently small to avoid constraints from BBN and CMB~\cite{Turner:1992tz}.
In addition, a technical concern could lead to a much larger tunneling rate.
At the origin of the SC potential,
the one loop perturbative calculation given in Eq.~\ref{CW} becomes no perturbative. In principle, the actual potential could be very different leading
to a much larger tunneling rate. The worry is unfounded.
Subsequently, a non-perturbative calculation~\cite{1997MPLA...12.2287L} has
been preformed which verifies that Eq.~\ref{CW} is still correct. Hence, the
Universe is safe from transitioning from the false vacuum to the true vacuum during inflation. In fact, the Universe never tunnels for even
arbitrarily low temperatures. For inflation to end, we will need to introduce new physics.
\subsection{ Ending Inflation (Witten Mechanism):}\label{witten}
Witten's mechanism generates a graceful exit and ends inflation.
We implement the Witten mechanism by introducing a set of QCD like fermions charged under a
non-Abelian gauge group with a Yukawa coupling to the SC field V$_{\text{y}}(\phi)=\sum y_i \phi\ast (q_R \bar q_L)_i+$ h.c.
where $y_i$ is the Yukawa coupling for
respectively the right and left-handed fermions $q_R$ and $q_L$.
The SC field $\phi$ is a singlet under the non-Abelian gauge group. Only some of the right handed quarks and left handed
quarks are charged under the U(1)$_X$ such that the Yukawa terms are gauge invariant. We also must
be careful how we assign U(1)$_X$ charges to the quark like fermions to ensure that the theory is anomaly free.
As with QCD, we can pick
a sufficiently small coupling constant such that the theory will only become strongly coupled at a low scale.
Finally, the fermion loop corrections have not been included in Eq.~\ref{CW}. Qualitatively the evolution of the aulos field
remains unchanged unless there are a large number of fermions with large Yukawa couplings.\footnote{
Witten used the same approximation and found a similar conclusion~\cite{Witten:1980ez}.}
When the temperature of the Universe drops below the strong coupling
scale $\Lambda_S$ of the non-Abelian gauge group, the vacuum
of the non-Abelian gauge group freezes, and the fermions gain a vev
\begin{equation}
\label{break}
\text{V}_y(\phi)\rightarrow\bigg(\sum y_i \langle q_R \bar q_L\rangle_i\bigg)\ast \phi+\text{h.c.}=\epsilon \ast\phi+\text{h.c.}.
\end{equation}
The vacuum seizes, which dynamically breaks any global symmetries
possessed by the quark like fermions $q_R q_L$, and the U(1)$_X$ gauge
symmetry.\footnote{
As Witten~\cite{Witten:1980ez} pointed out, a more rigorous analysis would replace
$\langle q_R\bar q_L\rangle(x)=$Q$(x)$ with an order parameter to describe symmetry breaking
in order to more carefully account for gauge invariance, but the underlying analysis would not change.
}
If some of the $y_i\simeq1$, then
$\epsilon\simeq\Lambda_S^3\simeq \text{T}_c^3$, where T$_c$ is the temperature at which the
vacuum of non-Abelian gauge group freezes. As we show below, the barrier trapping the SC field in the false vacuum goes away
once the non-Abelian gauge group freezes.
\vspace{10pt}
{\bf Concrete Example}
\vspace{10pt}
We now work through a concrete case of SCI with $\Lambda\simeq1$TeV and
$\langle|\phi|\rangle\simeq10$ TeV.
Initially, the Universe is radiation dominated. The Universe then cools and becomes stuck in the false vacuum.
The energy density of Universe becomes dominated by vacuum energy.
The Universe begins superluminal expansion once the temperature falls
below a few hundred GeV (N.B.\ The precise temperature when superluminal growth begins depends upon the
number of relativistic degrees of freedom. At temperatures between 100GeV to 1 TeV, there
are $\mathcal{O}(100)$ relativistic degrees of freedom
from the standard model. In which case, superluminal expansion begins once the temperature falls to $\sim300$ GeV).
As a conservative estimate, we will assume that inflation really only starts once the temperature of the Universe is a 100 GeV.
Inflation must last for at least 30 efolds to solve the horizon and flatness problems (See Eq.~\ref{efolds}).
After 30 efolds, the temperature of the
Universe drops to $\sim10^{-2}$~eV and the non-Abelian gauge group freezes. See Eq.~\ref{break}.
Once the non-Abelian gauge group freezes, the barrier goes away. Near the origin,
Eq.\ref{approx} together with Eq.\ref{break} has the form
\begin{equation}
(\epsilon\phi+ \text{h.c.})+\frac{1}{2}\,\tilde m^2|\phi|^2-\frac{1}{4}\,\tilde\lambda |\phi|^4
\end{equation}
(See Fig.\ \ref{apexinf}).
We note that the linear term will destabilize the meta-stable vacuum state ($\phi\simeq0$ ) by eliminating
the barrier if
\begin{equation}
\label{dest}
|\epsilon|\geq\frac{2}{3^{3/2}}\frac{\tilde m^3}{\tilde\lambda^{1/2}}.
\end{equation}
Upon substituting in values for $\epsilon$, $\tilde m$, and $\tilde\lambda$, we see that as the temperature drops so does m; while, $\tilde\lambda$
increases. In Fig.\ref{apexinf}, the barrier trapping the inflaton at the origin goes away once T $\simeq$ T$_c\sim10^{-2}$ eV,
with $g_X=0.4$,
M$_{\chi}\simeq$ 4 TeV and
T$_c\simeq\Lambda_S\simeq10^{-2}$ eV. In sum, the Universe goes through 30 efolds of inflation solving the
flatness and horizon problems, the barrier goes away, and
the SC field goes
rapidly to the true vacuum. The Universe then reheats as discussed below.
\subsection{Reheating}\label{reheating}
Once the barrier goes away, the field quickly goes to the
true minimum, where it may oscillate and decay into standard model particles. Reheating can be virtually
instantaneous for the decay rate $\Gamma_\phi \gg$H.
We might have a complicated decay process into standard model particles. For instance,
the SC field might decay into particles charged under U(1)$_X$ which subsequently decay into standard
model particles.
If the inflaton only partially decays into standard model particles,
then the SC sector might be related to dark matter. We will instead treat the case
where $\phi$ decays only into standard model particles and in particular Z bosons.
The choice is idiosyncratic; many other reasonable
decay routes exist (for instance, the SC field could decay into Higgs bosons).
In general, there are only 3 couplings between a
hidden sector and the standard model which are renormalizable and gauge invariant: a vector coupling
to hypercharge $B_{\mu \nu}$ (kinetic mixing), a scalar coupling to the Higgs HH$^\dagger$ ,
and a spinor coupling with a Higgs-neutrino operator HL~\cite{2009arXiv0909.4541W}. We will focus on the kinetic
mixing~\cite{Holdom:1985ag} between hypercharge $B_{\mu\nu}$ and the U(1)$_X$ SC gauge boson $X_{\mu\nu}$.
We have, thus, expanded the standard model $\rightarrow$ SU(3)$\times$SU(2)$_{\text{L}}\times$U(1)$_Y\times$U(1)$_X$.
In which case, the standard model will contribute to the CW potential Eq.~\ref{CW}. If the EW phase transition occurs during
inflation, then there will be a threshold correction to the CW potential.
The renormalization condition Eq.~\ref{CWcondition} can be
maintained by matching the renormalization group flow equations as it runs from the IR up to the threshold
and from the UV down to the threshold. The IR and UV are then necessarily sensitive to each other,
which only highlights the mysterious nature of the ``no bare mass" condition.
In the case of kinetic mixing between hypercharge U(1)$_Y$ $B^{\mu\nu}$
and the X boson U(1)$_X$ $X^{\mu\nu}$, the kinetic energy terms go like
\begin{equation}
\label{ke}
\mathcal{L}_{\text{KE}}^{BX}=-\frac{1}{4}B_{\mu\nu}B^{\mu\nu}-\frac{1}{4}X_{\mu\nu}X^{\mu\nu}+\frac{\chi}{2}B_{\mu\nu} X^{\mu\nu}
\end{equation}
where $\chi$ (at an effective level) is an arbitrary mixing parameter
and can take any value (which is how we treat $\chi$ for the remainder of the paper).
N.B.\ in some top down approaches the parameter $\chi$ can arise from integrating out vector like fermions
charged under both the hidden U(1)$_X$ and hypercharge~\cite{Holdom:1985ag}, which
gives $\chi\sim10^{-2}$.
Upon diagonalizing the kinetic term Eq.~\ref{ke} with a GL(2,R) rotation and then diagonalizing the gauge
boson mass matrix with an O(3) rotation, we can write $X_\mu$ in terms of the mass eigenstates of the standard
model Z boson and a Z$^\prime$~\cite{2009arXiv0909.4541W}. We now have the coupling of the SC field $\phi$ to the Z boson.
We can, then, determine the decay rate of the SC field into Z bosons.
When the mass of the
U(1)$_X$ gauge boson M$_{X}$ is large
compared to the unmixed Z boson mass M$_{Z_0}$ (the $0$ refers to the field before mixing)
or $\chi$ is small, the decay rate of the SC
inflaton into a pair of Z bosons~(\cite{Lee:1977yc}, ~\cite{Lee:1977eg}, and~\cite{Gunion:1989we}) is then
\begin{equation}
\label{Apexdecay}
\Gamma_{\phi\rightarrow \text{ZZ}}=\frac{\pi}{6} M_\phi (g_X\, \sin\theta_w \eta \Delta z)^4
\frac{\sqrt{1-x}}{x^2}(3x^2-4x+4)
\end{equation}
where M$_\phi$ ($\sim300$ GeV) is the mass of the SC inflaton, $\sin\theta_w$ is the Weinberg angle, $\Delta z=(M_{Z_0}/M_{X})^2$,
$x=4M_Z^2/M_\phi^2$, $\eta=\chi/\sqrt{1-\chi^2}$ and $g_X=0.4$. We find that
$\Gamma_{\phi\rightarrow \text{ZZ}}= 0.03$ eV$\gg$ H $\simeq10^{-4}$ eV,
where we have taken $\chi=.9$ and $M_{Z^\prime}\simeq M_X$ ($\sim4$ TeV). The field will predominantly decay into Z bosons,
if the masses of all other particles which couple to $\phi$ are more massive than $M_\phi$.
For instance the QCD like fermions have Yukawa couplings which are
$\mathcal{O}$(1) and have a mass $\sim$10 TeV.
We have assumed a large mixing $\chi$ between the $Z_0$ and $X$ boson but the large difference in the masses suppresses the decay rate.
One could imagine a slightly less massive X particle which would lead to a much larger decay rate. Regardless, the Universe rapidly reheats
converting the vacuum energy into radiation. With 100 relativistic degrees of freedom,
the Universe then reheats to a temperature $\simeq500$ GeV,
which is sufficiently high to do EW baryogenesis~\cite{Wainwright:2009mq}.
If we had a smaller mixing parameter, we could then have a smaller decay constant and
a lower reheat temperature.
The SC field can avoid collider constraints.
The Z$^\prime$ is heavy enough to avoid present collider bounds~\cite{2011PhRvD..83g5012A,Appelquist:2002mw}. The coupling
of $\phi$ to the standard model is strongly suppressed in this case due to large mass ratio of Z to Z$^\prime$.
Hence, the toy model is consistent with collider bounds. In a future paper, we will consider collider searches for the SC inflaton~\cite{doug1}.
\subsection{Scale of Inflation}\label{scale}
Cosmological considerations can place an upper and lower limit on the scale of SCI.
On the low end, any inflation model must satisfy Big Bang Nucleosynthesis (BBN) constraints.
In which case, the Universe then reheats to at least a few MeV.
Baryogenesis could potentially, also, place constraints on the minimal scale of inflation.
If the scale of inflation and reheat temperature
is above 100 GeV, then one can use EW baryogenesis or leptogenesis to generate a baryon
asymmetry~\cite{Wainwright:2009mq}. At present, we know of no published models of baryogenesis which work
with the scale of inflation below 100 GeV.
Thus, our present lack of imagination with regards to baryogenesis will require that inflation occurs above 100 GeV.
An aulos field might get around the above argument. As defined in the introduction, an aulos field is any field which has a small mass
during inflation and a large mass at the end of inflation. Later in the paper, we will use the aulos mechanism to generate density perturbations.
In a different direction, the aulos mechanism in conjunction with the Affleck-Dine Mechanism (ADM)~\cite{Affleck:1984fy}
could also be used to generate a baryon asymmetry even if the scale of inflation is below 100 GeV. We would like to emphasize
that the field generating a baryon asymmetry and the field generating density perturbations are not necessarily one and the same.
The Affleck-Dine Mechanism (ADM) can occur with a reheat temperature as low as a few MeV, but
ADM without the aulos mechanism requires that the scale of inflation be above $10^9$ GeV~\cite{Dine:2003ax,Banks:1996ea}.
The constraint on the scale of inflation from ADM
follows from 2 requirements: 1-- the mass of the field carrying baryon number must be larger than the mass of the proton
when the field decays. 2-- the field must also be Hubble-damped
during inflation.
An aulos field carrying baryon number can satisfy the 2 ADM requirements and could be used to do baryogenesis at a scale of inflation below
$10^9$ GeV.
The mass of the aulos field is small during inflation so the field is Hubble damped satisfying
the first requirement regardless of the inflationary scale. At the end of inflation, the mass of the field becomes large satisfying
the second requirement. The field begins to oscillate and generate baryon number via ADM.
In principle, one can have baryogenesis for a very low inflationary scale (potentially MeV scale).
We will leave detailed model building for a future paper~\cite{doug1}.
Regardless from observational constraints, the scale of inflation must still be larger than a few MeV.
On the other end of the scale, SCI most naturally occurs for $\Lambda\lesssim100$
TeV due to the appearance of non-renormalizable terms and
the number of efolds necessary to solve the horizon and flatness problems.
For instance, we have so far neglected a term which arises in general relativity
\begin{equation}
\label{gcoupling}
-\frac{1}{2} \xi R|\phi|^2,
\end{equation}
where $R=-32\pi G\Lambda^4$. $G$ is Newton's constant and $\Lambda^4$ is the vacuum energy during inflation. $\xi$
characterizes the coupling of the SC field $\phi$ to gravity. In the minimally coupled case ($\xi=0$),
there is no constraint on the scale of inflation. The scale can be arbitrarily large.
As another example, if the ``no bare mass" hypothesis is true, then it
is sensible to believe that the SC field is classically conformally invariant
which implies that $\xi=1/6$~\cite{Callan:1970ze}~\cite{Abbott:1981rg}.
We note that if we include loop corrections $\xi$ is a renormalizable coupling constant that runs to a fix point $\xi=1/6$ in the IR but
can have a negative value in the UV~\cite{Hill199221}.
Regardless, Eq.~\ref{gcoupling} can be problematic.
If $\xi>0$, Eq.~\ref{gcoupling} stabilizes the false vacuum, in which case inflation could never end.
If $\xi<0$, Eq.~\ref{gcoupling} could destabilize the false vacuum state, in which case inflation could end too soon.
Hence, we require that inflation ends before Eq.~\ref{gcoupling} becomes important
\begin{equation}
\label{gcrit}
\text{T}_c^2\gtrsim |\xi R|\propto |\xi| G \Lambda^4
\end{equation}
where T$_c$ is the critical temperature when inflation ends (See Eq.~\ref{dest}).
The number of efolds of inflation necessary to account for the horizon problem goes like
\cite{1990eaun.book.....K}
\begin{equation}
\label{efolds}
\text{N}_{min}=30 + \frac{1}{3} \left( 2 \log \left[ \frac{\Lambda}{1\text{ TeV}}\right]
+ \log\left[\frac{\text{T}_{RH}}{1\text{ TeV}}\right]
- \log\left[\gamma \right]\right)
\end{equation}
where $\Lambda$ is the scale of inflation, T$_{RH}$ is the reheat temperature at the end of inflation,
and $\gamma$ is the amount of
any subsequent entropy production after inflation and reheating.
Also, the number of efolds necessary to resolve the flatness problem is similar to the number
of efolds necessary to solve the horizon problem (See \cite{1990eaun.book.....K} Chapter on inflation).
We will require that the number of efolds of inflation be greater than or equal to Eq.~\ref{efolds}.
Tension between Eq.~\ref{gcrit} and Eq.~\ref{efolds} places an upper limit on the scale of SCI.
The necessary number of efolds increases with the scale of
inflation (first term of Eq.~\ref{efolds}), but Eq.~\ref{gcrit} pushes up T$_c$ with an increasing scale
of inflation.
We can only satisfy both Eq.~\ref{gcrit} and Eq.~\ref{efolds} if $\Lambda\lesssim100$ TeV for $\xi\sim\mathcal{O}(1)$.
SCI comfortably occurs between 100 TeV and 100 GeV (also potentially even the MeV scale),
but inflation at the TeV scale would certainly be
interesting. In many models beyond the standard model,
new machinery at the TeV scale is introduced to explain EW symmetry breaking:
gauge mediation, little Higgs models, large extra dimensions etc..
Theoretically, it is compelling to connect inflation to the dynamics which drive EW symmetry breaking.
Second, LHC and future colliders can effectively probe the TeV scale. Inflation could become a laboratory science.
At the moment, we can only probe inflation indirectly through cosmology.
Finally, the low scale of inflation inherently avoids many pitfalls of high scale inflation such as overclosure from gravitinos to moduli
and trans-Planckian physics.
\section{No Eternal Inflation}\label{eternal}
Virtually, all known models of inflation suffer from eternal inflation such as chaotic inflation~\cite{1986PhLB..175..395L} and new inflation
(See Fig.\ \ref{eternfig}).
There are a few exceptions such as~\cite{Baumann:2007np} which require a large amount of fine-tuning.
More generally as noted by Guth, most rolling models without eternal inflation
are pretty contrived from a field theoretical perspective~\cite{Guth:2000ka}.
As a simple example of eternal inflation consider chaotic inflation, which occurs at large field values.
A scalar field is displaced
from the true minimum. The potential of the scalar field V$(\phi)$ is sufficiently flat such that the field rolls very
slowly to the true vacuum state.
As the Universe rolls down, quantum fluctuations of the field
can cause a small patch of the Universe to go up the potential, which cause the fluctuation to grow (See Fig.\ \ref{eternfig}). A small patch can
continue to fluctuate up the potential until the quantum fluctuations of the field become order one i.e. $\Delta\rho/\rho \sim 1$,
at which point the patch can no longer roll back down. The patch begins to exponentially expand.
At which point, the Universe begins to inflate eternally and only an exponentially small fraction of the
Universe is not eternally inflating (See~\cite{Guth:2000ka} for an introduction to eternal inflation). These small pocket Universes
could evolve into a Universe which looks like ours or could be very different.
If we had a full picture of the eternally inflating Universe or Multiverse for short,
we could generate a probability distribution for the properties of different pocket Universes, such as
the cosmological constant, the amount of dark matter in the
Universe etc.. Then in fact, eternal inflation could be able to make various predictions and be tested.
Sadly, counting in an infinite Universe proves difficult.
To quote Alan Guth, ``In an eternally inflating Universe,
anything that can happen will happen; in fact, it will happen an infinite number of times.
Thus, the question of what is possible becomes trivial -- anything is possible,
unless it violates some absolute conservation law.
To extract predictions from the theory, we must therefore learn to
distinguish the probable from the improbable."~\cite{Guth:2000ka}.
Hence at worst, we have given up any predictive power of inflation
since anything and everything is possible, but if one had
a good way to count the relative occurrence of different pocket Universes,
we could regain the predictive power of inflation.
At present, there is no agreed method of determining the relative probability of different pocket Universes,
or in other words we have no sensible measure of the Multiverse.
In fact, eternal inflation and various measures so devised have instead led to a series of problems
such as the youngness problem, Boltzmann brains, etc., or make predictions which are wholly counterintuitive.
For instance,
the geometric or light cone measure (introduced by Bousso~\cite{Bousso:2006ev}) predicts
that time itself will end in the next 5 Billion years~\cite{Bousso:2010yn}.
(See \cite{Freivogel:2011eg} for a recent review). One might be interested in coming up with an inflationary model which avoids
eternal inflation in the first place.
\begin{figure}
\includegraphics[width=0.5\textwidth]{etern1.pdf}
\centering
\caption{Eternal inflation occurs at the origin of slowly rolling fields (as in new inflation)
and at large field values (as with chaotic inflation).
Quantum fluctuations prevent the field from rolling to the true vacuum as represented with the arrows.}
\label{eternfig}
\end{figure}
Eternal inflation occurs when
\begin{equation}
\label{EIcond}
\frac{\delta\rho}{\rho}\gsim1 \rightarrow \frac{V(\phi)^{3/2}}{m_{pl}^3}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} |V^{\prime}(\phi)|
\end{equation}
where $m_{pl}=2.4\times10^{18}$ GeV is the reduced Planck mass, $V(\phi)$ is the SC field but more generally is
any scalar potential. The prime refers to partial differentiation with respect to $\phi$.
For any scalar field, there are potentially two problematic field values.
First in the large field case ($\phi\rightarrow\infty$), the field could undergo chaotic eternal inflation as discussed above. Second near
the hilltop of the potential (a local maximum-- $V^\prime(\phi)=0$), a scalar field can also undergo eternal inflation as with the original new inflation models.
Regardless, SCI actually avoids eternal inflation. We consider both the hilltop and the chaotic cases.
In the hilltop case ($V^\prime(\phi)=0$), we show in the Appendix, that if
\begin{equation}
\label{NoEI}
\sqrt{|V^{\prime\prime}(\phi)|}>3\text{H},
\end{equation}
then there is no eternal inflation. Conversely if 3H$>\sqrt{|V^{\prime\prime}(\phi)|}$, then
there will be eternal inflation.
There is a hilltop point near the origin of the SC field ($\phi\simeq0$). We note that the tadpole term shifts the hilltop point away from the origin.
By plugging in numbers, it is clear
that there is no eternal inflation near the origin since slow roll is violated.
SCI can avoid chaotic or large field value eternal inflation.
In the large field case for the SC potential
Eq.~\ref{CW}, we violate Eq.~\ref{EIcond} once $\phi >m_{pl}$. We have so far not discussed the non-renormalizable
terms given in Eq.\ref{fp}. The SC potential $V(\phi)$ reaches a maximum before $\phi >m_{pl}$ by including a non-renormalizable term.
Non-renormalizable terms are a necessary component of an effective field theory.
For instance, a field which is non-minimally coupled to gravity has just such a term.
Non-minimal coupling was discussed in the previous section.
More generally, non-renormalizable terms can arise, when embedding a effective field theory in a more fundamental
theory. Non-renormalizable terms will go like $(\pm)\lambda_n \Lambda_c^{-n} \phi^{4+n}$, where $\Lambda_c$
is the ultra-violet cut off and $\lambda_n$ is positive.
We now include a non-renormalizable term in the potential of the SC field
Eq.~\ref{CW}
\begin{equation}
\label{CWredux}
V(\phi)=
\frac{\lambda}{4}|\phi|^4\bigg(\ln\bigg(\frac{|\phi|}{\langle\phi\rangle}\bigg)-\frac{1}{4}\bigg)
-\frac{\lambda_{2n}}{2n+4}\frac{|\phi|^{2n+4}}{\Lambda_c^{2n}}
\end{equation}
where $\lambda= 3g_X^4/8\pi^2$ (See Fig.\ \ref{noeternfig}).
For simplicity, we have neglected the cosmological tuning parameter in Eq.~\ref{CW}
(N.B.\ neglecting $\Lambda\ll\Lambda_c$ will not alter our results ) and the tadpole term which is irrelevant at large field values.
By inspection of Eq.~\ref{CWredux}, there exists a hilltop point $\phi_0$ where $V(\phi_0)$ is a maximum of the potential.
\begin{figure}
\includegraphics[width=0.5\textwidth]{noetern.pdf}
\centering
\caption{We have added a non-renomalizable term to the SC potential Eq.\ \ref{CW}.
Hence, the potential has no chaotic eternal inflation. In addition,
the potential does not have eternal inflation at the hilltops since the field does not slow roll at the hilltop points.}
\centering
\label{noeternfig}
\end{figure}
SCI also avoids eternal inflation at the hilltop point $\phi_0$.
Near $\phi_0$ we can approximate Eq.~\ref{CWredux} with
\begin{equation}
\label{CWsimple}
V(\phi)=\frac{\lambda^\prime}{4}|\phi|^4-\frac{\lambda_{2n}}{2n+4}\frac{|\phi|^{2n+4}}{\Lambda_c^{2n}}
\end{equation}
where $\lambda^\prime=\lambda \ln(\Lambda_c/\langle\phi\rangle)$ when
$4n^2 \ln(\Lambda_c/\langle\phi\rangle)>\ln(\lambda/\lambda_{2n})$
and $\ln(\Lambda_c/\langle\phi\rangle)>1/4$.
Then using Eq.~\ref{NoEI},
SCI will not have eternal inflation at $\phi_0$ if
\begin{equation}
\label{NoEIredux}
\frac{|\phi_0|}{\sqrt{2n+4}}<\frac{2}{3} m_{pl}\,\,\,\,\,\,\text{with}\,\,\,\,\,\,|\phi_0|=
\bigg(\frac{\lambda^\prime}{\lambda_{2n}}\bigg)^{\frac{1}{2n}} \Lambda_c,
\end{equation}
where $m_{pl}$ is the reduced Planck mass.
As an example, consider a dimension 6 operator and assume the cut off is the reduced Planck mass.
The non-renormalizable term is then
equivalent to Eq.~\ref{gcoupling} i.e.\ a non-minimal coupling to gravity.
We then find that there will be no eternal inflation for $\lambda_2>3/8\lambda^\prime\simeq10^{-2}$ or equivalent
in terms of Eq.~\ref{gcoupling} that $\xi< -0.1$. Also, we have avoided chaotic inflation since $|\phi_0|<m_{pl}$.
Hence, SCI
can avoid eternal inflation by introducing non-renormalizable terms or by simply coupling the SC field to gravity.
In fact, the above scenario is quite general.
$\xi$ is generally not constant and runs with the energy scale of the relevant interactions.
Hill \& Salopek~\cite{Hill199221} calculated the RGE of $\xi$ for a composite scalar field but they note that their conclusion holds for an
arbitrary scalar field as well. They found that $\xi=1/6$ is an IR fixed point.
In the UV (or at large field values), $\xi$ can easily become large and negative, which naturally
circumvents eternal inflation from happening as shown above.
In many ways it is not surprising that SCI avoids eternal inflation. Inflation driven by a rolling field requires an unusual situation with
the necessity of having an incredibly flat potential. After one has gone to the trouble of having an incredibly flat potential in the first place,
then eternal inflation can occur.
SCI does not depend upon slow roll dynamics and avoids eternal inflation.
\section{Density Perturbations}\label{perturbations}
The presence of a thermal bath (which is necessary in SuperCool (SC) inflation) will strongly
suppress curvature perturbations.
Adiabatic density perturbations go like
\begin{equation}
\label{curvpert}
\xi=\frac{\delta\rho}{\rho+p}
\end{equation}
where $\delta\rho$ is the variation of density and $\rho$
and $p$ are respectively the average energy density of the Universe and pressure at the time a perturbation
leaves the horizon. For a discussion on perturbations, see \cite{1990eaun.book.....K} and references there in i.e. \cite{Bardeen:1983qw}.
As with slow roll inflation, one might imagine introducing a single slowly
rolling scalar field $\phi_r$ to generate density perturbations. In this case, $\delta\rho\sim$ H V$^\prime(\phi_r)$
where H is the Hubble parameter when the perturbation leaves the horizon. V($\phi_r$) is the potential of the scalar field. In slow roll,
V$^\prime(\phi_r)=-3$H$\dot\phi_r$. The prime refers to partial differential with respect to $\phi_r$, and the
dot refers to differentiation with respect to time. $\rho+p$ is summed
over all matter and energy components.
Typically for inflation, $\rho+p=\dot\phi_r^2$, but if there is a thermal
background then we must include thermal radiation and pressure where $\rho_{r}=1/3p_{r}$. The thermal
component will dominate the sum $\rho+p$, which will suppress adiabatic perturbations generated by the rolling
field.
One might imagine that thermal variations might generate perturbations,
but the thermal variation across the initial patch is exponentially small. Equilibrium processes
wash out any temperature variations across the patch ($\delta \rho_{r}/\rho+p\ll1$), unless there is a process to generate
perturbations on all scales in the plasma just prior to the start of inflation.
We will not consider this case any further, but will leave the problem for a separate paper~\cite{doug1}.
Instead, we will turn to the aulos mechanism as discussed in the introduction.
\subsection{ Aulos Mechanism}\label{aulosmechanism}
We introduce an explicit model to implement the aulos mechanism as outlined in the introduction.
We introduce a new complex scalar field $\chi_a$ and 2 new scalar fields $\chi_b$ and $\chi_c$. The new fields have a potential
\begin{equation}
\label{toymodel}
\begin{aligned}
&V_{abc}=-\frac{\mu^2}{2}(\chi_b^2+\chi_c^2)+\frac{\lambda}{4}(\chi_b^2+\chi_c^2)^2\\
&+\frac{\lambda_\chi}{4}|\chi_a|^4\bigg(\ln\bigg(\frac{|\chi_a|}{\nu}\bigg)-\frac{1}{4}\bigg)+\sum_i^{abc}
\bigg(\frac{\alpha_i \text{T}^2}{2} |\chi_i|^2\bigg)\\
&+\frac{\lambda_b}{2}\bigg(\chi_b^2-\chi_c^2\bigg)|\chi_a|^2-\frac{\lambda_{a\phi}}{2}|\phi|^2 |\chi_a|^2.
\end{aligned}
\end{equation}
All of the coefficients are positive.
There is an O(2) symmetry between $\chi_b$ and $\chi_c$,
which is broken by the interaction term with $\chi_a$. $\chi_a$ has a U(1)$_a$ axial symmetry.
$\chi_b$ and $\chi_c$ get a vev
when ($\mu>\alpha_b$T$^2$ \& $\mu>\alpha_c$T$^2$).
\footnote{If inflation begins once the temperature of the Universe drops below 100 GeV, then we require
that $\mu_b$ and $\mu_c$ are $\geq$ 100 GeV
assuming $\alpha_i$ are order 1 for instance $\alpha=1/4$ for the SC field (See Eq.~\ref{approx}}).
We require that $\chi_a$ is independent of temperature.
If the field is temperature dependent then it will be difficult to generate the necessary
density perturbations for inflation. We could argue that $\chi_a$ has a zero temperature during inflation.
Instead, we cancel the temperature dependent term of
$\chi_a$ in Eq.~\ref{toymodel} ($\alpha_a$T$^2|\chi_a|^2$) by setting
\begin{equation}
\label{cancel}
\alpha_a= \frac{\lambda_b}{\lambda}(\alpha_c-\alpha_b),
\end{equation}
which requires a careful selection of the matter coupled to $\chi_a$, $\chi_b$, and $\chi_c$
given ($\lambda_b/\lambda$). The coefficients $\alpha_i$
in Eq.~\ref{toymodel} depends upon the matter content coupled to $\chi_i$.
We will require that $\chi_b$ and $\chi_c$ have different couplings to fermions i.e.\ $\alpha_c\neq\alpha_b$.
The different couplings to matter will then break the O(2) symmetry of Eq.~\ref{toymodel} and potentially
induce a different $\lambda$ for $\chi_b$ and $\chi_c$ due to quantum corrections.
One can insure that $\chi_b$ and $\chi_c$ do have the same $\lambda$
by again carefully choosing the matter content which couples to
$\chi_b$ and$ \chi_c$ such that the loop corrections will be the same for $\chi_b$ and $\chi_c$.
In addition, we will require that $\lambda_b/\lambda$ is a rational number to
allow for the cancellation of ($\alpha_a$T$^2|\chi_a|^2$). This will require a fair amount of model building.
In a more fundamental theory one might be
able to find such a scenario.
Otherwise naively, Eq.~\ref{cancel} requires severe tuning.
We do not address the naturalness of such a scenario at this time.
Regardless, the interaction term, now, cancels the $\alpha_a$T$^2\chi_a^2$ term;
$\chi_a$ has a Coleman Weinberg potential and then undergoes dimensional transmutation which spontaneously breaks
U(1)$_a$ i.e. ($\chi_a$ acquires a vev $\nu$).
With the vev of $\chi_a$ ($\nu\neq0$), we redefine $\chi_a$ in terms of a radial field ($\sigma$) and
an angular field ($a$) which corresponds to the
pseudo Nambu-Goldstone boson of $\chi_a=(\nu+\sigma)e^{i a/\nu}$.
We will associate the pseudo Nambu-Goldstone boson ($a$) with the aulos field. As one final note, we can avoid
eternal inflation in the aulos sector in much the same way that we avoided eternal inflation in the inflaton sector.
\subsection{ Aulos Mass} \label{aulosmass}
\begin{figure}
\includegraphics[width=0.5\textwidth]{aulos1.pdf}
\caption{Aulos potential Eq.~\ref{aulosPotential} in units of $\nu$. During inflation, the aulos vev and mass are small. At the end of inflation,
the vev and mass become large.}
\centering
\label{aulosfig}
\end{figure}
We introduce a mass term for the aulos field ($a$), by softly breaking the U(1)$_a$ symmetry. The potential of the aulos
field is then,
\begin{equation}
\label{aulosPotential}
\begin{aligned}
V(\chi_a)_S&=\frac{\lambda_s}{4}((\chi_a)^4+(\chi_a^\star)^4)\\
&=\frac{\lambda_s (\nu+\sigma)^4}{2}\bigg( \cos\bigg(\frac{4a}{\nu}+\pi\bigg)+1\bigg)
\end{aligned}
\end{equation}
where $\lambda_s$ is a dimensionless parameter which gives the size of the symmetry breaking
and $\chi_a^\star$ is the complex conjugate of $\chi_a$. A soft breaking term such as Eq.\ \ref{aulosPotential} can arise from a
constrained instanton first introduced by t'Hooft~\cite{PhysRevD.14.3432} and more fully developed by Affleck~\cite{Affleck:1980mp}.
We will leave a more careful analysis for the future and treat Eq.~\ref{aulosPotential} at an effective level.
In Eq.~\ref{aulosPotential}, we have shifted the aulos field such that
the minimum of the potential is given when $4a/\nu\rightarrow0$ (See Fig.~\ref{aulosfig}).
We might worry about the formation of domain walls. With the given breaking term
there are 4 degenerate vacua. But the worry is unfounded.
$\chi_a$ and the aulos field never reheat at the end of inflation.
In which case, the Kibble mechanism will not generate domain walls.
Expanding the cosine at the minimum gives that
\begin{equation}
\label{aulosMass}
m_a^2 =16 \lambda_s \nu^2=128\lambda_s f_a^2
\end{equation}
where $m_a$ is the mass of the aulos field and $f_a=\nu/4$ is the decay constant for the aulos field.
The mass is
proportionate to the spontaneous symmetry breaking scale.
Hence, the mass will vary if the spontaneously symmetry breaking scale changes.
\subsection{ Aulos Evolution}\label{aulsoevolution}
In our toy model during inflation, the aulos field is Hubble damped with 3H $\simeq10^{-3}$ eV $\geq m_a$.
The vev of the SC field is zero and does not contribute to the vev
of $\chi_a$. The vev of $\chi_a$ is initially small.
At the end of inflation, the SC field gets a large vev $\langle\phi\rangle$ which will generate a large and negative
effective mass term for $\chi_a$. Then from Eq.~\ref{toymodel}, $\chi_a$ gets a new large vev ($\nu$)
\begin{equation}
\label{vevafter}
\nu^2\rightarrow \frac{\lambda_{a\phi}\langle\phi\rangle^2}{\lambda_\chi}.
\end{equation}
The aulos mass and decay constant become large. See Eq.~\ref{aulosMass}.
When $m_a\geq$3H, the aulos field begins to oscillate and generates a condensate with an energy density
\begin{equation}
\label{aulosdensity}
\rho_a=\frac{1}{3}m_a^2 f_a^2 \theta_0^2=\frac{\lambda_s}{3}\nu^4 \theta_0^2
\end{equation}
where ($\theta_0=4a/\nu$) is the misalignment angle of the field when it begins to oscillate.
We assume that the field begins to oscillate near the top
of the cosine potential in Eq.\ref{aulosPotential}, in which case ($\theta_0\simeq\pi$).
We also evaluate $\rho_a$ with the enlarged values of $f_a$ and $m_a$. We have assumed
that the aulos field barely evolves as $f_a$ and $m_a$ grow to their maximum size.
At the moment, we have assumed that
when the aulos field begins to oscillate that $f_a$ is large and constant, but in fact, $f_a$ is not necessarily initially constant.
As the barrier trapping the SC field in the false vacuum goes away, it is unclear how the phase transition proceeds.
If the transition is first order and the SC field tunnels to the
true vacuum of the theory, then $f_a$ will in fact be constant. Otherwise,
the SC field will experience a period of rolling and potentially oscillating
around the minimum of the potential Eq.~\ref{CW}.
In this later case, if we want to determine precisely the
energy density in the aulos field and SC field, we need to solve
the coupled differential equations of motion for the aulos and SC field.
We leave this problem for a future paper. Regardless in our toy model,
we show that the decay rate of the SC field $\Gamma_\phi$ is
a factor of 10 larger than the decay rate for the aulos field $\Gamma_a$
and can be made much larger with a lighter Z$^\prime$ (See Eq.~\ref{Apexdecay}).
In addition, in the case of parametric resonance the inflaton will only oscillate once or twice before decaying.
We can in fact treat $f_a$ as constant during most of the evolution of the aulos field.
The aulos field in our model will predominantly decay into photons
(One could also consider a different decay route but leave that for future work).
We introduce a set of new chiral fermions charged
under the global U(1)$_a$ axial symmetry such that we can write down a Yukawa
coupling between the new chiral fermions and $\chi_a$. In addition, the new chiral fermions
are charged under EM, then due to the QED anomaly the aulos field can decay directly into photons with
\begin{equation}
\label{aulosDecay}
\Gamma_{a\rightarrow\gamma\gamma}=\frac{\alpha^{\prime2}}{64 \pi^3}\frac{m^3_a}{f_a^2}
\end{equation}
where $\alpha^\prime=g^{\prime2}/4\pi\times N^\prime$. $g^\prime$ is the coupling constant of the aulos fermions under U(1)$_{\text{EM}}$
and $N^\prime$ is a numerical factor which depends upon the number of fermions and their charge under EM.
We will take
$\alpha^\prime\simeq10^{-2} \alpha$ (which is reasonable if the fermions have a fractional charge under EM or couple due to kinetic mixing).
As long as the fermions are not charged under QCD and are heavy,
we can avoid constraints from rare decays and collider physics with $f_a\geq 100$ GeV and $m_a\geq1$ GeV (See \cite{Essig:2010gu} and
references within).
We also require the fermions to be sufficiently heavy to avoid LEP constraints.
The fermion's mass should be greater than roughly $100$ GeV.
When $\chi_a$ acquires a vev, the new chiral fermions become very massive TeV scale for $\mathcal{O}(1)$ Yukawa couplings.
\subsection{ Observational Signatures}\label{signatures}
Quantum fluctuations during inflation cause spatial variation of the misalignment angle $\theta_0$ of the aulos field
across the Universe. At the end of inflation, the aulos field begins to oscillate, once the mass of the aulos field
becomes larger than the Hubble parameter. At which point, variations in the misalignment angle during inflation
induce an isocurvature perturbation~\cite{Seckel:1985tj}.
Isocurvature are also
induced in the QCD axion~\cite{Seckel:1985tj} and the curvaton model but in a different manner from the aulos field.
The QCD axion and curvaton are
Hubble damped during inflation and experience fluctuations in their misalignment angle. Inflation ends; the Universe reheats and then cools.
Eventually, the Hubble parameter becomes smaller than the mass of the axion or curvaton.
The QCD axion and curvaton then begin to oscillate which generate isocurvature perturbations.
In contrast, the aulos field begins to oscillate and generate isocurvature perturbations immediately at the end of inflation
when the mass of the aulos field grows and becomes
larger than the Hubble parameter.
When the aulos field decays, the isocurvature perturbation
becomes a real curvature perturbation~\cite{Mollerach:1989hu}.
We can now give an expression for the normalization of the primordial power law spectrum~~\cite{Lyth:2002my}
\begin{equation}
\label{aulospert}
\Delta_{\xi}(k_0)= \frac{r\times q}{3\pi}\bigg( \frac{\text{H}}{f_a\theta_0}\bigg)\bigg\vert_{\text{Hor}}
\end{equation}
where $r=\rho_a/\rho$ is the ratio of energy density of the aulos field $\rho_a$ over the total energy density of the
Universe $\rho$ when the aulos field decays. The ratio of (H$/f_a \theta_0$) is to be evaluated during inflation, where H is the Hubble constant,
$f_a$ is the aulos decay constant, and $\theta_0$ is the misalignment angle of the aulos field.
We will take $q=1$,\footnote{$q$ parameterizes nonlinear effect in the
oscillation of the aulos field Eq.~\ref{aulosPotential}. See~\cite{Lyth:2001nq} \cite{Lyth:2002my} for more details.}
for the remainder of the paper.
The power-law index $n_s$ of the primordial power spectrum (following~\cite{Lyth:2002my}) is
\begin{equation}
\label{ns}
n_s\simeq1\pm\frac{2}{3}\frac{m_a^2}{\text{H}^2}+2\frac{\rho_r}{\Lambda^4}
\end{equation}
where respectively $(-)$ holds if during inflation the aulos field is near the top of the potential in Eq.~\ref{aulosPotential}
and conversely $(+)$ holds if the aulos field is near the bottom of the potential in Eq.~\ref{aulosPotential} (
WMAP 7 gives that $n_s=0.967$~\cite{Komatsu:2010fb}. We will assume that the aulos
field is stuck near the top of the potential and the $(-)$ case holds). In the second term of Eq.~\ref{ns},
$\rho_r$ is the energy density of radiation during inflation over the vacuum energy $\Lambda^4$.
During inflation the second term in Eq.~\ref{ns} can be safely ignored.
The aulos mechanism can also generate non-Gaussianities~\cite{Lyth:2002my} with
\begin{equation}
f_{\text{NL}}=\frac{5}{4r}.
\end{equation}
WMAP 7 constrains $f_{\text{NL}}\lesssim100$ which requires that $r\gtrsim10^{-2}$.
We will take $r= 3\times10^{-2}$, which gives an $f_{\text{NL}}=42$. We can reasonably
have a much smaller $f_{NL}$ but picked a value which would be observable.
In the near term, Planck and 21-cm experiment can detect ($f_{\text{NL}}\simeq5$ to 10)~\cite{PhysRevLett.107.131304}.
In addition, the aulos field can isocurvature perturbations. At present, the size of isocurvature perturbations must be less
than about a tenth of the size of adiabatic density perturbations~\cite{Komatsu:2010fb}.
The isocurvature constraints for the aulos field are the same as with the curvaton
(See \cite{Lyth:2002my} for more details). We will just summarize the effects. Isocurvature constraints require baryogenesis and
dark matter (DM) genesis
to occur after the decay of the aulos. In which case, there will be no isocurvature perturbations.
Instead, if the aulos field can decay directly into baryons or DM, then the aulos field could generate interesting levels of
isocurvature perturbations, which future experiments could detect.
We will leave the last possibility for future research and will only require that the aulos field decays before
the production of DM and baryons. Below, we show that the aulos field for the parameters chosen decays almost instantaneously.
At that time,
the temperature of the
Universe is around 500 GeV. Hence, one can use EW baryogenesis to generate the baryon asymmetry,
which occurs once the temperature of the Universe is around
100 GeV~\cite{Wainwright:2009mq}.
As to dark matter, we leave that for a separate paper, but the QCD axion
would be a viable DM candidate.
\subsection{ Aulos Parameters-Colliders}\label{collider}
We can now relate the parameters of the aulos field to the primordial power spectrum.
The size of the soft breaking term of U(1)$_a$ (See Eq.~\ref{aulosPotential})
\begin{equation}
\label{lambda_s}
\lambda_s= 6\pi^2 (1-n_s)\bigg(\frac{\Delta_{\xi}(k_0)\,\theta_0}{r \times q}\bigg)^2.
\end{equation}
can be given in terms of $n_s$, $r$, $q$ and ($\theta_0\simeq\pi$). $\Delta_{\xi}^2(k_0)=2.43\times10^{-9}$~\cite{Komatsu:2010fb}.
We find that $\lambda_s=4.3\times10^{-5}$, which is small but technically natural.
With inflation at the TeV scale and $r=3\times10^{-2}$, Eq.~\ref{aulospert} fixes $\nu\simeq4\times10^{-2}$ eV
during inflation. Also, $\lambda_{a\phi}\lesssim1$, such that the aulos field does not inadvertently destabilize the SC field before
inflation ends. The small vev $\nu$ generates an effective negative mass term for the SC field. Conversely if $\lambda_{a\phi}\sim1$,
we could then use the small negative mass term to end SCI rather than the Witten mechanism.
This negative mass case would be similar in spirit to thermal inflation. We will leave this case for future research.
At the end of inflation, $\nu=7$ TeV and $m_a=730$ GeV with $\lambda_\chi\simeq\lambda_{a\phi}$
and $\langle\phi\rangle\simeq 10$ TeV.
For simplicity, we will assume that the vacuum energy associated with
the $\chi_i$ fields is subdominant compared to the scale of inflation ($V_{abc}\lesssim\Lambda^4$),
which implies that $\lambda_{a\phi}\lesssim10^{-4}$.
In which case, the mass of $\chi_a$ is still heavier than $M_X$. Then, $\phi$ will not decay into $\chi_a$.
One could imagine a more complicated decay process, but we will ignore the possibility at the moment.
Eq.~\ref{aulosDecay} gives the decay rate
of the aulos field ($\Gamma_a=2\times10^{-3}$ eV, with $\alpha^\prime=10^{-2} \alpha$),
which is long compared to the decay rate of the SC field ($\Gamma_\phi=0.03$ eV)
and short compared to the Hubble scale at the end of inflation ( H$=4\times10^{-4}$eV).
So in fact, the aulos and SC field decay almost instantaneously relative to H.
The aulos field can make connections between collider physics and the CMB.
We can redefine the mass and the decay rate of the aulos field in terms of parameters
determined by cosmological measurements i.e.\ Eq.\ref{lambda_s} and the coupling of aulos field to fermions.
SC vev and the aulos decay constant should be on the same scale. Due to the low scale of
SCI, colliders could rule out the scenario or potentially give a route of discovery.
In addition to LHC,
there will be new searches for hidden sectors at Jefferson Lab, JPARC, and future Long Baseline Neutrino experiments
such as LBNE~\cite{Essig:2010xa,Essig:2010gu}. We will leave a detailed discussion of particle physics searches for SCI
to a future paper~\cite{doug1}.
\section{Conclusion}\label{conclusion}
In conclusion, we have proposed SuperCool (SC) inflation as a solution to old inflation The Universe
inflates solving the horizon and flatness problems. At a sufficiently low temperature, a non-Abelian gauge group becomes strongly
coupled which generates a tadpole term.
The new term destabilizes the false vacuum, providing a graceful exit.
Hence, SuperCool Inflation (SCI) generates a successful period of inflation.
There have been other models proposed which have a graceful exit,
such as double field inflation~\cite{Adams:1991ma}, a variant of hybrid inflation~\cite{PhysRevD.49.748},
and chain inflation~\cite{Freese:2004vs}~\cite{Freese:2006fk} to name a few.
These models with the exception of chain inflation use the dynamics of a rolling field to give old inflation a graceful exit.
The rolling models will have eternal inflation and must confront the measure problem unlike SCI.
Furthermore, the rolling models are finely tuned to generate a sufficient amount of inflation~\cite{Adams:1991ma}.
In SuperCool (SC) inflation, the running of a non-Abelian gauge coupling determines the number of efolds of inflation. Thus, SCI
is no more finely tuned than QCD, with the caveat that the ``no bare mass" condition is sensible.
SC inflation is not the only inflationary model, which takes advantage of
thermal effects such as warm inflation~\cite{Berera:1995ie} and thermal inflation~\cite{Lyth:1995ka}.
Warm inflation
modifies the friction term for the equation of motion of the inflaton due to the appearance of a finite temperature during inflation
and should be lumped in with other rolling models.
Thermal inflation and SC inflation are very similar in spirit. In thermal inflation, an effective mass
from the finite temperature potential as with SCI stabilizes a false vacuum.
Eventually a negative mass term destabilizes the false vacuum and the field transitions to the true vacuum.
Unlike SC inflation, thermal inflation is too brief
to solve the flatness and horizon problems and does not have a way to generate density perturbations
but is sufficiently long to dilute unwanted relics such as gravitinos etc..
As argued in the text, the TeV scale seems to be a natural fit to SC inflation.
The scale of SC inflation could be much lower than the TeV scale with a hard lower bound given by BBN ( a few MeV).
In principle, there is no upper bound on the scale of SC inflation if one can ignore non-minimal coupling to
gravity or non-renormalizable terms in general. But as argued in the text, one should not neglect for instance a non-minimal coupling
to gravity. In which case, there is an upper bound on the scale of SC inflation on the order of 100 TeV.
TeV scale inflation prompts a question on the naturalness of the initial conditions. We have assumed as our starting
point that the initial patch which hosts the visible Universe today was hot, homogeneous, and isotropic when inflation began.
We do not address the naturalness of such a set up, but the same ambiguity existed in Guth's original model.
In that instance, Guth argued that quantum gravity just might generate the right initial conditions. Even in rolling models
one still faces the same difficulty or one might hope that eternal inflation offers a route out of initial condition conundrum,
but then one must address the measure problem.
At the moment, there is no obvious solution to the initial condition problem in any model of inflation, without invoking
eternal inflation, which is good or bad depending upon your tastes.
As an interesting feature, SC inflation naturally avoids eternal inflation. As argued in the text, eternal inflation appears both ubiquitous and
fraught with difficulty. One solution to eternal inflation is no eternal inflation. Eternal inflation occurs at hilltop points (the maximum of the potential)
and at large field values (chaotic inflation). See Fig.\ \ref{eternfig}. Non-minimal coupling to gravity of the
SC field or non-renormalizable
terms in the effective potential for the SC inflaton can prevent chaotic eternal inflation. At hilltop points,
the field has no slow roll. We show in the Appendix that the field will then not have eternal inflation at the hilltop points.
In fact, we should not be surprised
that eternal inflation does not occur for SC inflation. Eternal inflation for a rolling field is a symptom of slow roll which generally
requires a large amount of tuning~\cite{Adams:1990pn}.
Only once we have gone to the trouble to finely tune our potential to get slow roll that we run into eternal
inflation. SC inflation does not depend upon slow roll. Hence, SC inflation avoids eternal inflation.
One of the great success of slow roll inflation is generating adiabatic density perturbations. The SC inflaton generates
the vacuum energy which drives inflation and ends inflation due to the Witten mechanism, but does not generate
appreciable density perturbations.
We have imagined two possible ways to generate perturbations. First, if there exist initial perturbations in the plasma just
before inflation begins, then those perturbations on the largest scales\footnote{
We only observe perturbations on the largest scales which correspond to perturbations that leave the horizon
within the first 10 efoldings of inflation. On small scales equilibrium processes wash out perturbations, but
large scale perturbations can persist.
}
will redshift as the Universe inflates and produce
the adiabatic density perturbations we see today. For instance, one might imagine
a first order phase transition just before inflation begins. Bubble collisions might
seed perturbations in the plasma of just the sort we need.
In this paper we have not addressed this scenario and leave such considerations to a separate paper~\cite{doug1}.
Instead we introduced a new mechanism to generate
density perturbations which we have dubbed the aulos mechanism.
The aulos field is a pseudo Nambu-Goldstone boson similar to an axion.
De Sitter fluctuations induce variations in the misalignment angle of the aulos field. At
the end of inflation the mass of the aulos field becomes large.
At which point, the aulos field begins to oscillate and generate a condensate.
The variation in the misalignment angle induces an isocurvature perturbation. Real adiabatic density perturbations
are generated when the aulos field decays into radiation. As noted previously, the aulos mechanism successfully
matches observations from the CMB and potentially
can generate non-Gaussianities and isocurvature perturbations, which could be observed by Planck.
We define an aulos field more broadly as any
pseudo Nambu-Goldstone boson which has a small spontaneous symmetry breaking scale and mass during inflation
and a large mass and breaking scale at the end of inflation.
An aulos field has several interesting applications to inflation.
We only mention a few.
In a manner similar to the curvaton scenario, the aulos mechanism could be used to generate
perturbations in an arbitrary inflation scenario.
In a separate publication~\cite{doug1}, we use the aulos mechanism to generate perturbations in hybrid models at the TeV scale,
which avoids previous constraints \cite{Lyth:1999ty}.
The curvaton scenario requires that the scale of inflation to be above $10^{10}$ GeV.
The aulos mechanism works for an arbitrarily low scale of inflation.
Second, the Affleck-Dine mechanism typically requires that the scale of inflation be near $10^{9}$ GeV.
An aulos field which carries baryon number could do baryogenesis via an Affleck-Dine
mechanism for an arbitrary scale of inflation.
One could potentially implement inflation successfully at the MeV scale.
At the moment, we know of no models of inflation, which is successful
with the scale of inflation below 100 GeV!
We can imagine that the aulos mechanism can be used beyond inflationary dynamics.
The aulos mechanism can be used to generate
spatial variation of $\alpha$, which has been observed by~\cite{2010arXiv1008.3907W} and a spatial variation of dark energy.
In addition, the aulos mechanism could generate compensated perturbations,
which has recently garnered interest. See \cite{2010ApJ...716..907H,2011arXiv1107.1716G} for a discussion.
In the last two cases, the aulos field would not be associated with a field which generates density perturbations during inflation.
We will consider these possibilities in separate publications.
\bigskip
{\it Acknowledgements}: We would like to thank P. Adshead, B. Bardeen, S. Dodelson, E. Kolb, J. Frieman,
R. Feldmann, P. Fox, N. Gnedin, A. Guth, D. Hooper, J. Kopp, W. Hu, J. Lykken, G. Miller, E. Neil, A. Stebbins, A. Upadhye,
E. Weinberg, M. Wyman and A. Zablocki for discussion and feedback.
We would like to extend a special thanks to M. Turner and C. Hill whose criticism profoundly reshaped the final paper. Finally,
A. Albrecht's encouragement, criticism, and insight was a light in the dark.
This material is based upon work supported in part by the National Science
Foundation under Grant No. 1066293 and the hospitality of the Aspen Center for Physics.
D. Spolyar is supported by the Department of Energy.
\bibliographystyle{ieeetr}
|
\section{Introduction}
\label{section:intro}
A successful theory of modified gravity has proved hard to find. Most theories begin by postulating new physical principles, for example, the existence of fundamental fields with certain symmetry properties \cite{Nicolis_Galileons} or additional dimensions \cite{Rubakov, MaartensKoyama}; see \cite{MG_report, Nojiri_Odintsov_review} for comprehensive surveys of the current literature. By introducing such principles on which to base a theory we are immediately selecting specific directions in theory space to investigate. An alternative approach is to cautiously explore outwards from the corner of theory space that we understand best, that is, General Relativity (GR). One way to implement this strategy is to construct a parameterized framework that allows for small deviations from GR in a cosmological context, akin to the Parameterized Post-Newtonian framework that has been used to test GR extensively within the Solar System \cite{Will1971, Thorne_Will, Will_Nordvedt_1972, Will2011}.
In order to extract maximum benefit from this approach we need to develop a sense of how the parameters we use impact the growth of structure.
Historically, our theories of structure formation have been developed using matter-dominated cosmological models \cite{Landau, Peebles, Peebles_book, Blumenthal, Meszaros}. Though not appropriate for realistic calculations, a toy Einstein-de Sitter (EdS) cosmology remains an immensely useful testing ground for new theories; here, we can gain a qualitative understanding of perturbation evolution whilst our knowledge of the background expansion remains on a firm footing. Therefore, maintaining the principle of caution advertised above, we will consider the effects of parameterized gravity on an EdS universe (which contains pressureless matter only). This will be an excellent approximation to the matter epoch of the real universe, and has the added advantage that analytic solutions are achievable in some cases due to the simple properties of pressureless matter.
Despite the wide variety of modified gravity theories present in the literature, a survey of field equations reveals some common features. In many theories the gravitational constant appearing in the Poisson equation acquires a time-dependence and/or scale-dependence \cite{WiggleZ_Nesseris}, and the Newtonian gravitational potential $\Psi$ and curvature perturbation $\Phi$ (in the conformal Newtonian gauge) are not equal as they are in GR. Parameterized frameworks for modified gravity are usually constructed to incorporate such properties. We can ask what typical effects these generic features might have on observables: for example, in addition to affecting the growth rate of structure, distortion of gravitational potentials will imprint secondary anisotropies on the CMB. However, we will assume that modifications to GR must be negligible at very early times, to avoid significantly impacting the primary CMB and the sensitive reaction rates of Big Bang Nucleosynthesis \cite{Bambi_Giannotti,Calabrese,SPT_EDE}; see \cite{Brax_Davis} for consideration of scalar-field models that are non-negligible at the time of recombination.
This paper investigates the evolution of cosmological perturbations in the parameterized framework implemented in \cite{Skordis2009, Ferreira_Skordis, Baker2011, Zuntz2011}, for an EdS background.
\textsection\ref{section:param_choice} introduces the necessary formalism in the context of theories that are constructed purely from metric quantities. In \textsection\ref{section:delta} we use this framework to calculate the evolution of density perturbations on intermediate and large scales. Two other quantities of interest are also calculated: the growth function $f(z)$, and the Integrated Sachs-Wolfe effect that is induced in such gravity theories (note that this is not the same as the late-time Integrated Sachs-Wolfe effect that occurs in a $\Lambda$-dominated era).
In \textsection\ref{section:extra_dof} we take the first steps towards implementing a similar treatment of gravitational theories that introduce additional scalar degrees of freedom. New degrees of freedom cause a significant increase in complexity which renders a general, model-independent calculation almost impossible, at least analytically. We will implement a parameterized effective fluid approach known as `Generalized Dark Matter' (GDM) \cite{HU_GDM} to facilitate the treatment of these additional scalars, and consider the case in which the effective fluid has a negligible equation of state. This approximation excludes some classes of theories from our analysis, such as $f(R)$ gravity, but is applicable in other cases; we drop the restriction again in \textsection\ref{section:zeta} . In \textsection\ref{section:zeta} we consider a perturbation that is conserved on super-horizon scales in GR, and ask under what conditions this fact remains true in parameterized modified gravity. The conclusions of this work are presented in \textsection\ref{section:discussion}.
The purpose of this paper is to develop an understanding of parameterized gravity, not to pursue accurate calculations for the real universe.
Hence the plots and trial parameter values used here are intended to be illustrative rather than realistic.
\section{Parameterization of Metric-Only Gravity Theories}
\label{section:param_choice}
No standard parameterization of modified gravity currently exists. A common choice is to introduce a free function that describes any time- or scale-variation of the gravitational constant in the Poisson equation \cite{Bean, DanielLinder,Pogosian_parameterization,Hojjati_Zhao_2011,Bertschinger_Zukin,Hu_Sawicki}:
\begin{equation}
-2 k^2\Phi=\kappa a^2 \mu (a,k)\,\rho\Delta\label{Poisson}
\end{equation}
where $\kappa=8\pi G_0$, $G_0$ is the canonical value of Newton's gravitational constant, \mbox{$\Delta=\delta+3{\cal H}\theta$} is a comoving gauge-invariant density perturbation and $\theta$ is the velocity potential of a fluid defined by \mbox{$v_i=\nabla_i\theta$}.
Often a second free function is used to describe the ratio of the two conformal Newtonian potentials, $\eta (a,k)=\Phi/\Psi$. By redefining \mbox{$\zeta=(1-1/\eta)$} this can be rewritten:
\begin{equation}
\Phi-\Psi=\zeta (a,k)\Phi\label{ph_slip}
\end{equation}
The arguments of the `modified gravity functions' (MGFs) $\mu$ and $\zeta$ will be suppressed hereafter. We will call this form of the slip relation parameterization B. In \cite{Baker2011} it was argued that a parameterization of this type is only applicable in the quasistatic regime, and that in the case of purely metric theories it implicitly corresponds to higher-derivative theories. This is true if one assumes that eqn.(\ref{ph_slip}) is an exact `template' for the slip relation of a modified gravity theory, which means that $\zeta$ can only be a function of homogeneous background quantities. If instead one is prepared to let $\zeta$ be a function of non-homogeneous environmental variables and initial conditions then the slip relation may not necessarily imply a higher-derivative theory; it becomes impossible to ascertain the derivative order of the theories being parameterized without further information \cite{private_comm}.
In \cite{Baker2011} an alternative format was suggested, in which the order of derivatives in the field equations is made explicit. In this alternative parameterization a metric theory with a $\Lambda$CDM background incurs extra constraint equations (see Table \ref{tab:2nd_order_constraints} in the appendix), which force the slip relation to be \cite{Skordis2009, Ferreira_Skordis}:
\begin{equation}
\Phi-\Psi=\zeta\Phi+\frac{(\mu-1)}{{\cal H}\mu}\dot\Phi
\label{sf_slip}
\end{equation}
This slip relation will be termed parameterization A. The key feature to note here is that in parameterization A the MGF $\mu$ appears in both the Poisson and slip equations, unlike parameterization B. This apparently small difference leads to some degree of ambiguity in the interpretation of current constraints on the MGFs \cite{Zuntz2011}.
However, the purpose of this paper is not to discuss the subtleties of parameterization choice at length; we wish to keep our results as general as possible. We can treat both of the parameterizations simultaneously by adopting the following Poisson and slip equations:
\begin{eqnarray}
-2 k^2\Phi&=&\kappa a^2 \mu_P \,\rho_M\Delta_M\label{gen_Poisson}\\
\Phi-\Psi&=&\zeta\Phi+\frac{\mu_s-1}{{\cal H}\mu_s}\dot\Phi\label{gen_slip}
\end{eqnarray}
To recover parameterization B we set $\mu_s=1$ but keep $\mu_P$ general. To recover parameterization A we set $\mu_s=\mu_P$. Note that $\mu_s$ is related to the function $\tilde{g}$ in \cite{Baker2011} by \mbox{$\mu_s=(1-\tilde{g})^{-1}$}. Throughout this section we will sometimes leave results expressed in terms of the three functions $\mu_P,\,\mu_s$ and $\zeta$. We wish to emphasize from the outset that there are only really two independent functions, and all expressions should be evaluated in either the A-type or B-type instance. We write our expressions in this general format because can be instructive to see whether the modified terms have their origin in the Poisson equation (indicated by the presence of $\mu_P$) or the slip relation (indicated by $\mu_s$ and $\zeta$).
What theories map onto equations (\ref{gen_Poisson}) and (\ref{gen_slip})? The answer is `very few', which is a cause for concern given that the above forms are often used to obtain constraints on modified gravity from current data. To map \textit{exactly} onto these parameterizations a theory must stem from an action that is constructed only from curvature invariants and leads to fields equations that contain at most second- or third-order time derivatives (for parameterizations A and B respectively). By the immense power of Lovelock's theorem \cite{Lovelock1, Lovelock2} such a theory can only differ from GR if it introduces either nonlocality or spacetimes of dimension greater than four. This is a very restricted class of theories, though examples do exist \cite{Deser_Woodard, Deffayet_Woodard, Zumino}.
The limitations described above can be relaxed if one is satisfied with an approximate correspondence between theories and parameterization, rather than an exact one. For example, it is frequently assumed that the perturbed Einstein equations will retain the form of eqns.(\ref{Poisson}) and (\ref{ph_slip}) in theories with additional scalar degrees of freedom. This cannot be exactly true; any new scalar coupled to gravity will modify the zeroth-order Einstein equations in some way, and we expect to see perturbations of the new field appearing in the linearized Einstein equations. However, the form of eqns.(\ref{Poisson}) and (\ref{ph_slip}) is retained within a limited range of distance scales for some theories \cite{Baker2011,Amendola_WL,Schimd,Pogosian_Silvestri,KoyamaMaartens_structureDGP}.
To avoid such approximations we will proceed by taking eqns.(\ref{gen_Poisson}) and (\ref{gen_slip}) at `face value', i.e. assuming that there are no new scalar degrees of freedom hidden behind them. This is the assumption that is implicitly being made if equations such as (\ref{Poisson}) and (\ref{ph_slip}) are implemented in an Einstein-Boltzmann solver \cite{MGCAMB,Hojjati_Pogosian,Hojjati_Zhao_2011,Dossett_2011} and used to generate ISW and matter power spectra.
In \textsection\ref{section:extra_dof} we will introduce an extended parameterization that attempts to account for the additional scalars explicitly.
\section{Density Perturbations}
\label{section:delta}
In this section we will consider how the growth of cold dark matter (CDM) density perturbations in an EdS universe is influenced by the MGFs. This is a model for the matter-dominated epoch of the real universe. The growth of structure during an epoch in which dark energy is also relevant was investigated in \cite{Ferreira_Skordis}, using a specific ansatz for $\mu_P$ and $\zeta$.
We will assume that any time-variation of the MGFs during the matter-dominated epoch must be very small in order to prevent them from evolving to a region of parameter space that would cause conflict with observations. This is not guaranteed, for example, in \cite{Zuntz2011} we found that cancellations between MGFs can lead to observables very close to the predictions of $\Lambda$CDM. Hence even very radical departures from GR can be accommodated by current data in finely-tuned situations; however, for the purposes of this paper we will assume that our present universe does not correspond to such a case. We will therefore take the time derivatives of $\mu_P, \mu_s$ and $\zeta$ to be negligible in comparison to the rate of evolution of other variables. We intend to relax this restriction in future work.
Although we are neglecting their time-dependence, $\mu_P$ and $\zeta$ may still contain scale-dependence. They are dimensionless functions, but scale-dependence can appear as a ratio to some special scale that arises in a given theory, i.e. \mbox{$k/k_*$}. An example of such a privileged scale is the Compton wavelength of the scalaron in $f(R)$ gravity \cite{Sotiriou_fR}.
The fluid conservation equations for CDM energy density and momentum perturbations are given by:
\begin{eqnarray}
\dot\delta_M&=&-k^2\theta_M+3\dot\Phi\label{CDM_delta}\label{fluid_eqn1}\\
\dot\theta_M&=&-{\cal H}\theta_M+\Psi\label{CDM_theta}\label{fluid_eqn2}
\end{eqnarray}
where $\theta_M$ is the velocity potential. Differentiating eqn.(\ref{CDM_delta}) and combining with eqn.(\ref{CDM_theta}) leads to a second-order equation for $\delta_M$:
\begin{equation}
\label{delta_evol_1}
\ddot\delta_M+{\cal H}\dot\delta_M - 3\ddot\Phi-3{\cal H}\dot\Phi+k^2\Psi=0
\end{equation}
This is the same as found in GR. However, differences from GR arise when we use a non-trivial slip relation to eliminate $\Psi$.
Substituting eqn.(\ref{gen_slip}) into eqn.(\ref{delta_evol_1}):
\begin{equation}
\hspace{-0.4cm}
\label{delta_evol_2}
\ddot\delta_M+{\cal H}\dot\delta_M - 3\ddot\Phi-3{\cal H}\dot\Phi\left[1+\frac{k^2}{3{\cal H}^2}\left(\frac{\mu_s-1}{\mu_s}\right)\right]+k^2(1-\zeta)\Phi=0
\end{equation}
Before studying the behaviour of this equation it is useful to delineate a hierarchy of distance scales
\begin{enumerate}
\item The nonlinear scale on which clusters and galaxies form.
\item The quasistatic scale on which time derivatives of perturbations can be neglected in comparison to their spatial derivatives.
\item Larger scales on which the above approximation is no longer valid, but are still well within the horizon.
\item Scales that are greater than our observable horizon.
\end{enumerate}
We will consider the solutions of eqn.(\ref{delta_evol_2}) in regions 3 and 4. In fact it is possible to derive a single equation for $\Phi$ that is valid in both regions 3 and 4, and then use the Poisson equation to relate its solutions to $\delta_M$ (we thank C. Skordis for pointing this out). We will use an equivalent method that is simpler but a little less elegant.
\subsection{Subhorizon scales}
\label{sub:subhorizon_scales}
In region 3 described above we can approximate \mbox{${\cal H}/k \ll 1$} and \mbox{$\Delta_M\approx\delta_M$} since \mbox{$|\theta_M|\sim |v_M|/|k|$} is small. We use derivatives of eqn.(\ref{gen_Poisson}) to eliminate $\Phi$ from eqn.(\ref{delta_evol_2}):
\begin{eqnarray}
&&\ddot{\delta}_M+{\cal H}\dot\delta_M\left[1+\frac{3}{2}\mu_P\frac{(\mu_s-1)}{\mu_s}\right]\\
&&\hspace{2.0cm}-\frac{3}{2}{\cal H}^2\mu_P\delta_M\left[1-\zeta+\frac{(\mu_s-1)}{\mu_s}\right]=0 \nonumber
\label{delta_evol_3}
\end{eqnarray}
where we have used the result $\dot\ensuremath{{\cal H}}=-\frac{1}{2}\ensuremath{{\cal H}}^2$ in an EdS universe. Writing the solutions of this equation in the form
\begin{equation}
\delta_M = N^{+} \,a^{\frac{n^{+}}{2}}+N^{-}\,a^{\frac{n^{-}}{2}}
\end{equation}
where $N^{+}$ and $N^{-}$ are constants and the power-law indices are:
\begin{eqnarray}
\label{n_general}
n^{\pm}&=&-\frac{1}{2}\left(1+3\mu_P\frac{(\mu_s-1)}{\mu_s}\right)\\
&&\pm\frac{1}{2}\left[9\frac{\mu_P^2}{\mu_s^2}(\mu_s-1)^2-30\frac{\mu_P}{\mu_s}+6\mu_P(9-4\zeta)+1\right]^{\frac{1}{2}} \nonumber
\end{eqnarray}
In parameterization A this reduces to
\begin{equation}
n^{\pm}_{A}=\left(1-\frac{3}{2}\mu\right)\pm\frac{1}{2}\sqrt{9\mu^2+12\mu(3-2\zeta)-20}\label{n_SF}
\end{equation}
whereas in parameterization B it becomes
\begin{equation}
n^{\pm}_{B}=-\frac{1}{2}\pm\frac{1}{2}\sqrt{1+24\mu_P(1-\zeta)}\label{n_phenom}
\end{equation}
It can be verified that in the limit $\mu_s=\mu_P=1$ and $\zeta=0$ eqn.(\ref{n_general}) recovers the GR result $\delta_M \propto a$. We can see immediately that in both parameterizations there is a term $-24\mu_P\zeta$ that leads to degeneracy between the effects of the individual MGFs. Note that $\zeta$ only appears within this degenerate combination, so it cannot significantly impact growth if $\mu_P$ is small. We note that if a $\mu$-like MGF is implemented in the Poisson equation for $\Psi$ instead of $\Phi$ then this degeneracy does not arise in the parameterization B case \cite{Pogosian_Silvestri, Hojjati_Pogosian,Hojjati_Zhao_2011}.
When $n^{\pm}$ are imaginary the solutions for $\delta_M$ are damped oscillations. However, since our calculation has neglected the effects of baryons or radiation this oscillatory behaviour simply indicates unphysical solutions rather than anything meaningful. To have at least one growing mode we need $n^{+}$ to be positive, for which the relevant condition is:
\begin{equation}
\label{growing_bound}
\mu_s (2-\zeta) > 1
\end{equation}
We expect that an approximate version of this bound should be obeyed in the real universe, in order to reproduce the observed matter power spectrum - see \textsection\ref{sub:constraints}. However, in the real universe the hard bound of eqn.(\ref{growing_bound}) will be softened by contributions to growth from the radiation and $\Lambda$-dominated eras. Note that there is no restriction that prevents $\zeta$ from adopting negative values.
We wish to understand the physical mechanisms through which the MGFs are exerting their influence on small scales. We can get a feel for this by thinking about eqn.(\ref{delta_evol_3}) in the context of a simple mechanical system. The last term on the righthand side represents a time-dependent forcing that drives the collapse of density perturbations. The $\dot{\delta}_M$ term is analogous to a frictional force, which in familiar physical situations always acts to oppose motion; its magnitude decreases with time due to the factor of $\ensuremath{{\cal H}}$. Since we are parameterizing around a $\Lambda$CDM background the evolution of $\ensuremath{{\cal H}}$ is unaffected by the MGFs, and hence cannot be contributing to deviations from GR.
The overall magnitude of the driving term is controlled by $\mu_P$. This intuitively makes sense -- if we increase the gravitational coupling strength then we expect structures to collapse faster. Less intuitive is the appearance of $\mu_s$ and $\zeta$ in the driving term, which have the ability to change its sign. We will assume throughout that \mbox{$\mu_P,\, \mu_s>0$} always to maintain agreement with our physical notion of attractive gravity, but note that there is no such restriction on $\zeta$. The condition for the driving force to maintain the same direction as it has in GR is exactly eqn.(\ref{growing_bound}). It is interesting to see that in parameterization A $\mu_s$ and $\zeta$ can have counteracting effects on the driving term.
Qualitatively, a negative $\zeta$-value with large magnitude enables one to weaken $\mu_P$ considerably whilst maintaining growth during a matter-dominated epoch. If parameterization B is adopted this effect does not exist because the modification to the Poisson equation has no influence on the sign of the driving term.
The friction term has a somewhat simpler behaviour, as it is unaffected by $\zeta$. In parameterization B the friction is unchanged from GR, but in parameterization A \mbox{$\mu=\mu_P=\mu_s$} has the power to enhance or suppress friction effects. A value of $\mu>1$ acts to increase friction, but simultaneously strengthens the driving term that drives perturbations to collapse. One expects that some degree of cancellation between these two effects may be possible, even without the extra freedom provided by $\zeta$.
\subsection{Superhorizon scales}
\label{sub:superhorizon_scales}
On large scales we ought not to make the approximation \mbox{$\Delta_M\approx \delta_M$}, as the magnitude of the velocity potential is not negligible. We will adopt a different strategy, using the linearized Friedmann equation to solve for $\Phi$ as a proxy for $\delta_M$.
We first adopt a phenomenological approach (in the spirit of parameterization B), and assume that Newton's constant is identically modified in all perturbed field equations. We then have (neglecting a small term proportional to $k^2$) :
\begin{equation}
-{\cal H}(\dot\Phi+{\cal H}\Psi)=\frac{\kappa a^2}{6} \rho_M\mu_P\delta_M=\frac{{\cal H}^2}{2}\mu_P\delta_M\\
\end{equation}
Differentiating, and using that $\dot\delta_M\approx 3\dot\Phi$ on large scales (from eqn.(\ref{fluid_eqn1})):
\begin{equation}
\ddot{\Phi}+\frac{\cal H}{2}(1+3\mu_P)\dot\Phi+{\cal H}\dot\Psi=0
\end{equation}
Using the slip relation (eqn.(\ref{gen_slip})):
\begin{equation}
\ddot\Phi+\dot\Phi\frac{\cal H}{2}\left[3+3\mu_P-2\zeta\right]=0
\end{equation}
The power-law solutions are $\Phi\propto a^{p^{\pm}}$, with $p^{+}=0$ and \mbox{$p^{-}=(2\zeta-2-3\mu_P)$}. On these scales $\delta_M$ follows the behaviour of $\Phi$ up to a constant offset, which can be set to zero by initial conditions. So the dominant mode -- constant potential outside the horizon and $\delta_M$ frozen -- is the same as in GR. However, the decaying mode is affected by the MGFs. For $\zeta=0$, increasing $\mu_P$ will result in faster decay. This seems somewhat counterintuitive, since one would usually associate an increase in gravitational strength with \textit{reduced} decay of overdensities. However, since we are working on superhorizon scales gauge issues may invalidate our physical notions of gravitational growth and decay.
In parameterization A the linearized Friedmann equation is
\begin{equation}
-6{\cal H}(\dot\Phi+{\cal H}\Psi)=\kappa a^2 \rho_M\delta_M+A_0 k^2\Phi \label{00E_SF}
\end{equation}
where
\begin{eqnarray}
A_0&=&-2\left(\frac{\mu_P-1}{\mu_P}\right)\left(1+\frac{{\cal H}^2}{Q}\right) +2\zeta\frac{{\cal H}^2}{Q}\nonumber\\
\mathrm{with}\quad\;Q&=&{\cal H}^2+\frac{k^2}{3}-\dot{\cal H}\nonumber
\end{eqnarray}
The derivation of the above expression is described in appendix \ref{app:Z_consv}.
Note that $\mu_P$ does not feature explicitly in eqn.(\ref{00E_SF}). Newton's constant is not modified directly -- instead one considers all possible additional terms that could appear in the linearized Einstein equations, which can be determined up to a dimensionless function of background quantities . In the case of second-order metric-only theory the only terms that can be added to the linearized Friedmann and `0i' equations are proportional to $\Phi$, which can be absorbed into an effective Newton's constant.
On very large scales $\lim_{k\to 0}(k^2A_0)=0$, so we can neglect the extra term on the RHS of eq.(\ref{00E_SF}). Repeating the steps we took for the parameterization B-like case we obtain $p^+=0, \,p^-=\mu_P\left(2\zeta-5\right)$. The values of $p^{-}$ in the two parameterizations converge as one tends to the GR limit, as of course they must.
However, the decaying mode is not hugely interesting as it is unobservable (unless there are some very radical modifications to GR involved). The important result here is that the potential remains constant on superhorizon scales, as usual.
\subsection{Connection to Constraints}
\label{sub:constraints}
To what extent are our results for an idealized EdS model borne out in the real universe? Fig.\ref{constraints_plot} shows the joint constraints on the MGFs $\mu_P$ and $\zeta$ obtained using the following data sets: the 7-year WMAP CMB data \cite{wmap7}, the SDSS DR7 matter power spectrum \cite{sdss-dr7}, a prior $H_0 = 73.8 \pm 2.4$ \cite{riess2011}, the BBN constraint $\Omega_b h^2 = 0.022 \pm 0.002$ \cite{bbn}, and the Union2 Supernova Ia data \cite{Amanullah2010}. The differences between the two sets of contours were discussed in detail in \cite{Zuntz2011}. Here our main interest is the extent to which the constraints reflect the analytic solutions of the previous two subsections.
\begin{figure}[t!]
\begin{center}
\includegraphics[width=9.2cm]{both2.png}
\caption{Joint constraints on the slip parameter $\zeta$ and $\mu_P$ at $z=0$, for parameterization A (black lines) and B (filled green). In both cases 68\% and 95\% contours are shown.}
\label{constraints_plot}
\end{center}
\end{figure}
Given the restriction that $\mu_s$ must be positive, eqn.(\ref{growing_bound}) tells us to expect that $\zeta < 2$ on subhorizon scales in parameterization A. For $\zeta=0$ we must have $\mu_P=\mu_s>1/2$ for density perturbations to grow during the matter era, which corresponds to the approximate location of the near-vertical black contour. This contour is not the result of any artificially-imposed boundary, but delineates a very sharp fall-off in the likelihood distribution for $\mu_P$ (see figure 6 of \cite{Zuntz2011}). In contrast, in parameterization B any $\mu_P>0$ permits growing modes for $\zeta=0$, giving rise to the more gradual fall-off shown by the shaded countours.
However, eqn.(\ref{growing_bound}) also implies that in parameterization B $\zeta<1$ is necessary for growth of CDM density perturbations on subhorizon scales, which is contradicted by Fig.\ref{constraints_plot}. This is not too surprising -- we expect the simple bounds implied by our EdS example to be blurred by the complexities of a realistic cosmological model. If models in the region of parameter space $\zeta>1$ experience sufficient growth during radiation- and $\Lambda$-dominated epochs, or on scales outside the validity of eqn.(\ref{n_phenom}), then they will not be excluded by an MCMC analysis.
The degeneracy between $\mu_P$ and $\zeta$ is visible in both contour sets, but it is more pronounced in parameterization B. This is because in parameterization A the quadratic term in eqn.(\ref{n_SF}) makes it more difficult to accommodate the effects of a large $\mu_P$ with a small $\zeta$, or vice-versa.
\subsection{Other Growth Observables}
\label{sub:growth_observables}
\subsubsection*{Integrated Sachs-Wolfe Effect}
\label{subsub:ISW}
A cosmological model governed by GR will not experience an Integrated Sachs-Wolfe (ISW) effect \cite{Sachs_Wolfe} during a matter-dominated epoch. The kernel of interest for the ISW effect is $\dot\Phi+\dot\Psi$ (in the conformal Newtonian gauge), which in GR is equal to $2\dot\Phi$. From the standard Poisson equation one has, on scales well below the horizon:
\begin{equation}
\Phi\propto a^2\rho_M\delta_M
\end{equation}
Energy-momentum conservation gives $\rho_M\propto a^{-3}$, whilst in GR $\delta_M\propto a$, leaving $\Phi$ with zero time-dependence. The situation only changes for $z \lesssim 0.5$ when $\Lambda$ begins to suppress the growth rate of $\delta_M$ \cite{Crittenden_Turok}.
In modified gravity this behaviour is affected in two ways: a non-trivial slip relation will cause the ISW kernel to differ from $2\dot\Phi$, and the scaling of $\delta_M$ with $a$ will be altered. Then generically we expect a non-zero ISW effect, even during a matter-dominated phase of the universe \cite{Zhang_2006, Giannantonio_2010}. We will calculate the contribution to the CMB temperature power spectrum of this effect for the metric theories discussed in \textsection\ref{section:delta}.
We begin by using the slip relation (eqn.(\ref{gen_slip})) to express the ISW kernel purely in terms of $\Phi$:
\begin{equation}
\label{ISW1}
\dot\Phi+\dot\Psi=\frac{1}{2}\dot\Phi\left[3-2\zeta+\frac{1}{\mu_s}\right]-\ddot\Phi\left(\frac{\mu_s-1}{{\cal H}\mu_s}\right)
\end{equation}
Using the Poisson equation to connect $\Phi$ and $\delta_M$, we have (discarding the decaying mode):
\begin{eqnarray}
\Phi&=&M(k)\eta^{(n^{+}-2)}\nonumber\\
\mathrm{where}\;\;M(k)&=&-\frac{\kappa \delta_{M,0}(k)\mu_P \rho_{M,0}}{2k^2}\nonumber
\end{eqnarray}
$n^{+}$ is given by eqn.(\ref{n_general}) and density perturbations are normalised by their present values i.e. \mbox{$\delta_M (k,\eta)=\delta_{M,0}(k) \,\eta^{n^{+}}$}. Generally one expects the growth rate of density perturbations to be scale-dependent in a modified gravity theory, whereas in GR all linear subhorizon modes grow at the same rate. However, we are modelling a region of theory space close to GR and hence we will assume negligible variation of the growth rate over the range of $k$ relevant to observations of the ISW plateau.
So the ISW kernel is:
\begin{eqnarray}
\dot\Phi+\dot\Psi&=&\frac{M(k)}{2}(n^{+}-2)\eta^{(n^{+}-3)}\times\nonumber\\
&&\left[6-n^{+}-2\zeta+\frac{1}{\mu_s}(n^{+}-2)\right]\label{ISW2}
\end{eqnarray}
Note that in the GR limit $n^{+}=2$ the above expression vanishes as expected.
Next we need to compute the power spectrum of this modified-gravity induced ISW effect. The expression to be evaluated is:
\begin{equation}
C_l=\frac{2}{\pi}\int_0^{\infty}dk\,k^2P(k)\,\Big| \frac{\Theta_l(k)}{\delta_{M,0} (k)}\Big|^2
\label{general_Cl}
\end{equation}
The temperature perturbation observed today, $\Theta_l(k, \eta_0)$, consists of three parts:
\begin{eqnarray}
\Theta (k,\eta_0)&=&\:\mathrm{monopole\;term}\:+\:\mathrm{dipole\;term}\:\\
&&\hspace{-0.5cm}+\int^{\eta_0}_0d\eta\:e^{-\tau}\left[\dot\Phi (k,\eta)+\dot\Psi(k,\eta)\right] j_l[k(\eta_0-\eta)] \nonumber
\end{eqnarray}
Let us focus on the dominant contribution to the ISW power spectrum which comes from the $(\dot\Phi+\dot\Psi)^2$ term, and ignore the cross terms with the monopole and dipole. The cross-terms should only yield small corrections because the monopole and dipole terms are evaluated at the time of last scattering, and hence affect different $l$-values from the subsequent ISW effect. We will take the visibility function $e^{-\tau}$ to be a step function at recombination. Of course, if we are considering a truly EdS universe then there is no recombination event or time of last scattering, but this detail is irrelevant -- we are only interested in modelling the real universe well after recombination. In the real universe recombination occurs sufficiently early that for practical purposes we can take $\eta_{\mathrm{rec}}\approx 0$.
A standard derivation relates the power spectrum of density fluctuations today to the power spectrum of the primordial potential \cite{Dodelson}. The only modification to this that occurs in our theory is a factor of $\mu_P^{-2}$, which arises when we relate density fluctuations to $\Phi$ via the Poisson equation. However, this is cancelled by a factor of $\mu_P^2$ in $M(k)$. The scales of interest to us are sufficiently large that we can set the transfer function $T(k)\sim 1$. Then, for a Harrison-Zel'dovich spectrum we find
\begin{eqnarray}
\label{Cl_ISW}
\hspace{-0.2cm}C_l^{ISW,sq}&=&\frac{9\pi}{4}(n^{+}-2)^2 \left(6-n^{+}-2\zeta+\frac{1}{\mu_s}(n^{+}-2)\right)^2\delta_H^2\nonumber\\
&&\hspace{-5mm}\times\int_0^{\infty}\frac{dk}{k}\left[\int^{\eta_0}_0\,\eta^{(n^{+}-3)}j_l[k(\eta_0-\eta)]\,d\eta\right]^2
\end{eqnarray}
where $\delta_H$ is the amplitude of primordial perturbations at horizon-crossing during a single-field slow-roll inflation scenario.
The superscript `$sq$' on $C_l$ reminds us that we have only evaluated the ISW-squared term and not the cross-terms.
A plot of this power spectrum for several combinations of $\mu$ and $\zeta$ in parameterization A is shown in Fig.\ref{fig:ISW}. The normalization of the y-axis is arbitrary because we have not attempted an accurate calculation of the $C_l$s; we are more interested in how the shape and amplitude of the power spectrum is affected by different parameters. Note that our subhorizon solution for $\Phi$ is likely to become invalid for the largest scales, hence the region $l\lesssim 10$ in Fig.\ref{fig:ISW} is not fully accurate.
\begin{figure}
\includegraphics[scale=0.5]{output_plot_final.png}
\caption{Low-$l$ power spectrum of the ISW effect induced by a theory of modified gravity constructed in parameterization A (see \textsection\ref{section:param_choice} for details). The y-axis scaling is arbitrary. The region $l\lesssim 10$ is expected to be subject to corrections.}
\label{fig:ISW}
\end{figure}
Provided that $\mu_P$ is not unusually small, $\zeta$ can have a significant effect on the amplitude of the power spectrum. For example, compare the curves with $\{\mu_P,\zeta\}$ equal to $\{1,-0.25\}$ and $\{2,0\}$ -- a small change in $\zeta$ dominates over a large change in $\mu_P$.
The models $\{2.0,-0.25\}$ and $\{3.0,0.0\}$ have the same value of $n^+$, and hence the same spectral shape. It may seem a little surprising that a model with parameters $\{1.0,-0.25\}$ predicts a larger ISW effect than one with $\{2.0,-0.25\}$ for $l>6$; this is because a larger value of $n^+$ shifts the dominant contribution to the integral in eqn.(\ref{Cl_ISW}) to later times, therefore shifting the corresponding power spectrum left towards larger scales.
In general, however, it is difficult to cleanly disentangle the effects of $\mu_P$ and $\zeta$ on the ISW power spectrum, because they appear in a degenerate combination inside $n^+$.
\subsubsection*{Growth Function}
\label{subsub:growth_function}
The rate of growth of structure as a function of redshift is often quantified via the \textit{growth function}:
\begin{equation}
\label{growth_function}
f(z)=\frac{d\,\mathrm{ln}\,\Delta_M}{d\,\mathrm{ln}\,a}
\end{equation}
In GR $f(z)$ is independent of the wavenumber $k$, but in a modified gravity scenario this is not generally the case. However, if we assume we are not dealing with very radical departures from GR then this scale-dependence is likely to be small over a restricted range of $k$. On the scales of region 3 (see the beginning of this section) where $\Delta_M\sim\delta_M$, the growth function is simply \mbox{$f(z)=\frac{n^{+}}{2}$}, where $n^{+}$ is given by eqn.(\ref{n_general}).
\vspace{1mm}
\section{Theories With Additional Degrees of Freedom}
\label{section:extra_dof}
Many gravity theories introduce degrees of freedom (d.o.f.) other than perturbations of the metric -- additional scalar or vector fields, second metrics, or any combination thereof \cite{MG_report}. For model-independent constraints to be genuinely feasible we need to construct a parameterization that is able to accommodate such d.o.f. In this paper we will restrict ourselves to scalar d.o.f. only, which find widespread motivation from particle physics, braneworld models and string theory.
In \textsection\ref{section:param_choice} the parameterization-A-based approach was to add to each linearized Einstein equation all possible terms that could appear in the context of a metric-based second-order theory, which amounted to terms in $\Phi$ and $\dot\Phi$ multiplied by some function of background quantities (see appendix \ref{app:Z_consv}). When we extend this framework to include extra d.o.f. two new types of terms appear. Firstly we must allow for
perturbations of the d.o.f themselves -- for example, in a Brans-Dicke theory \cite{BransDicke} one expects the perturbations of the scalar field $\delta\phi,\,\delta\dot{\phi}$ and $\delta\ddot{\phi}$. Such new scalars are awkward to work with directly without knowing their underlying equations of motion. Therefore we will adopt an alternative approach based on a scheme by Hu \cite{HU_GDM} and treat this first type of additional term as perturbations of an effective fluid. A similar approach was adopted in \cite{Koivisto_Mota, Mota_2007}.
In a fully general case one should allow the effective fluid to have a time-varying equation of state, non-adiabatic perturbations and significant anisotropic stress. This would render our simple EdS model invalid by modifying the evolution of the cosmological background. To proceed with our analytic treatment we will sacrifice some generality by setting the equation of state of the effective fluid to zero. It then contributes to the zeroth-order Friedmann equation and supports density and velocity perturbations, but has negligible pressure and anisotropic stress.
This approximation is not unreasonable; for example, in Ho\v{r}ava-Lifshitz gravity \cite{HL1, HL2} a non-local Hamiltonian constraint gives rise to an integration constant which may be interpreted as a pressureless fluid. The additions to the Einstein equations in linear Einstein-Aether theory also behave as a pressureless fluid during an EdS phase, as can be seen using eqns.(15) and (16) of \cite{Zlosnik_structure_growth}. The effective fluid of Eddington-Born-Infeld gravity can experience a CDM-like phase too \cite{Banados_Ferreira_Skordis}. However, making this restriction does exclude some important classes of theories, such as $f(R)$ gravity, except for special choices of $f(R)$.
The second type of new term that arises when extra d.o.f. are included is the following combination of metric potentials:
\begin{equation}
\label{gamma}
\Gamma=\frac{1}{k}\left(\dot\Phi+\ensuremath{{\cal H}}\Psi\right)
\end{equation}
In this paper we are working in the conformal Newtonian gauge, but the parameterization we are using is constructed using gauge-invariant quantities -- see \cite{Skordis} for details. In the gauge-invariant formalism $\Phi$ and $\Psi$ are replaced by the Bardeen potentials $-\Psi_H$ and $\Phi_A$. If the background equations are fixed to be $\Lambda$CDM, as in \textsection\ref{section:delta}, then the gauge-invariant version of $\Gamma$ cannot appear in the linearized field equations because it introduces higher-order time derivatives of the scale factor. However, when new d.o.f. are added these unwanted time derivatives can be cancelled by perturbation of the new scalar. This somewhat technical point was demonstrated explicitly for scalar-tensor theories in appendix B of \cite{Baker2011}.
Concretely, then, we write the Einstein equation in the form:
\begin{equation}
\label{einstein}
G_{\mu\nu} \;=\; 8\pi G_0 a^2\, T_{\mu\nu}+a^2 U_{\mu\nu}
\end{equation}
where the tensor $U_{\mu\nu}$ contains all the non-standard terms arising from a theory of modified gravity. At the perturbed level the components of $\delta U_{\mu\nu}$ are then given by (in Fourier space):
\begin{eqnarray}
\label{2nd_order_Us}
U_{\Delta} &=& A_0 k^2\Phi+F_0k^2\Gamma+\kappa a^2 \rho_E\delta_E\nonumber\\
U_{\Theta}&=& B_0 k\Phi+I_0k\Gamma+\kappa a^2 \rho_E\theta_E\nonumber \\
U_P &=& C_0 k^2\Phi + C_1 k\dot{\Phi}+J_0 k^2\Gamma+J_1 k \dot\Gamma \nonumber\\
U_{\Sigma} &=& D_0\Phi+\frac{D_1}{k} \dot{\Phi}+K_0\Gamma+\frac{K_1}{k} \dot\Gamma \label{U_2nd_order}
\end{eqnarray}
where
\begin{eqnarray}
\label{U_cmpt_def}
U_{\Delta}&=&-a^2\delta U^0_0, \qquad \vec{\nabla}_i U_{\Theta}=-a^2 \delta U^0_i \nonumber\\
U_P&=&a^2\delta U^i_i, \qquad\,\, D_{ij}U_{\Sigma}=a^2(\delta U^i_j-\frac{1}{3}\delta U^k_k\delta^i_j)
\end{eqnarray}
The coefficients $A_0,\dots K_1$ above are functions of background quantities, dependencies which we will suppress to avoid cluttered expressions.
A subscript $E$ denotes the quantities relating to the effective fluid. The expressions above contain fewer terms than the corresponding ones in \cite{Baker2011}; the additional metric terms present in that paper are those needed to form a gauge-invariant combination with perturbations of the new scalar. Here we have kept those terms folded into the effective fluid so that the correspondence with $\delta U_{\mu\nu}$ in a general gauge is clearer.
The system to be solved then comprises of the six variables \{$\Phi, \Gamma, \delta_E, \theta_E, \delta_M,\theta_M$\} and six dynamical equations: the two spatial components of the linearized Einstein equation and two fluid conservation equations for each of CDM and the effective fluid (equivalent to Bianchi identities). Note that because $\delta U_{\mu\nu}$ contains additional metric perturbations the conservation equations of the effective fluid will contain some non-standard terms. The full system of equations is displayed in appendix \ref{app:dof_eqns}.
$\Gamma$ can be eliminated from the two spatial Einstein equations to give a second-order equation in terms of $\Phi$:
\begin{eqnarray}
\label{dof_phi_eqn}
\ddot\Phi&+&\dot\Phi\left[\frac{\dot\alpha}{\alpha}-\frac{\dot{W_1}}{W_1}+k\frac{W_2}{\alpha}+\frac{kZ_1}{\alpha}\right]\\
&&+\frac{k}{\alpha}\Phi\left[\dot{W_2}-\frac{\dot{W}_1W_2}{W_1}+k\frac{W_2 Z_1}{\alpha}-k\frac{W_1Z_2}{\alpha}\right]=0\nonumber
\end{eqnarray}
where $\ensuremath{{\cal H}}_k=\ensuremath{{\cal H}}/k$ and
\begin{eqnarray}
\label{alpha_etc_array}
\alpha&=&\left(D_1-\frac{1}{\ensuremath{{\cal H}}_k}\right)(J_1-6)+K_1\left(9\ensuremath{{\cal H}}_k-C_1+\frac{2}{\ensuremath{{\cal H}}_k}\right)\nonumber\\
W_1&=&K_1\left(3\ensuremath{{\cal H}}_k-J_0-\frac{2}{\ensuremath{{\cal H}}_k}\right)+(J_1-6)\left(K_0+\frac{1}{\ensuremath{{\cal H}}_k}\right)\nonumber\\
W_2&=&(J_1-6)(D_0-1)+K_1(2-C_0)\nonumber\\
Z_1&=&\left(\frac{1}{\ensuremath{{\cal H}}_k}-D_1\right)\left(3\ensuremath{{\cal H}}_k-J_0-\frac{2}{\ensuremath{{\cal H}}_k}\right)\nonumber\\
&&+\left(\frac{1}{\ensuremath{{\cal H}}_k}+K_0\right)\left(9\ensuremath{{\cal H}}_k-C_1+\frac{2}{\ensuremath{{\cal H}}_k}\right)\nonumber\\
Z_2&=&(2-C_0)\left(\frac{1}{\ensuremath{{\cal H}}_k}-D_1\right)+(D_0-1)\left(9\ensuremath{{\cal H}}_k-C_1+\frac{2}{\ensuremath{{\cal H}}_k}\right) \nonumber
\end{eqnarray}
We have assumed that time derivatives of the MGFs are negligible, consistent with our treatment of metric-only theories. If we set the MGFs to zero the third term vanishes and we recover the GR result $\Phi=$ constant. Modifications to the Poisson equation mean that the solution for $\Phi$ is not as easily translated into a solution for $\delta_M$ as it is in GR. One route is to substitute the solution for $\Phi$ into either of eqns.(\ref{Einstein3}) or (\ref{Einstein4}) and solve for $\Gamma$, then use both of these solutions in eqn.(\ref{delta_evol_1}) -- see appendix \ref{app:deltas_in_dof} for an explicit calculation.
We wish to consider eqn.(\ref{dof_phi_eqn}) on superhorizon and subhorizon scales, as we did in \textsection\ref{section:delta}. However, there is some uncertainty involved in taking this limit without knowing specifically the functional forms hiding behind the MGFs. We will assume that any scale-dependence appears relative to some preferred scale of a given theory, i.e. as a function of $(k/k_*)$.
\subsection{Subhorizon Scales}
\label{subsub:dof_subhorizon}
Under the assumptions stated above, and retaining only the dominant terms when expanded in powers of $(k/\ensuremath{{\cal H}})$, on subhorizon scales eqn.(\ref{dof_phi_eqn}) reduces to:
\begin{equation}
\ddot\Phi+\frac{(2D_1-C_1+2K_0-J_0)}{(6-J_1+2K_1)}k\dot\Phi+\frac{(2D_0-C_0)}{(6-J_1+2K_1)}k^2\Phi=0 \label{dof_phi_ss}
\end{equation}
With the change of variable $x=k\eta$ the above equation can be rewritten (using primes to denote derivatives with respect to $x$):
\begin{eqnarray}
\label{dof_phi_ss_2}
&&\Phi^{\prime\prime}+\frac{\beta}{\gamma}\Phi^{\prime}+\frac{(2D_0-C_0)}{\gamma}\Phi=0\\
&&\mathrm{where}\;\beta=(2D_1-C_1+2K_0-J_0),\;\;\gamma=6-J_1+2K_1.\nonumber
\end{eqnarray}
Without taking specific forms of the MGFs we cannot say whether the coefficients of the second and third terms above are positive or negative, only that sufficiently large modifications to gravity have the power to flip their signs (although the magnitude of such non-GR terms is expected to be small in the domain of validity of this equation). However, one might expect \mbox{$(2D_0-C_0)/\gamma>0$} and \mbox{$\beta/\gamma>0$} so that changes in $\Phi$ damp out (and therefore return to the GR-like situation) rather than grow.
Eqn.(\ref{dof_phi_ss_2}) has the form of a simple mechanical oscillator, and will display the usual phenomena of ringing or over-damping depending on the values of the coefficients. Specifically, its behaviour will depend on the value of \mbox{$\beta^2/\gamma-8D_0+4C_0$}, with negative values of this quantity leading to damped oscillations in $\Phi(x)$ and positive values leading to exponentially growing and decaying solutions.
Let us apply the single boundary condition that the potential is constant on superhorizon scales; we will see shortly that this is likely to remain true in theories with extra d.o.f. We can then determine the subhorizon solution up to an overall constant:
\begin{eqnarray}
\Phi(x)&\propto&e^{m_+x}-\frac{m_+}{m_-}e^{(m_+-m_-)}e^{m_-x}\\
\mathrm{where}\;\;m_{\pm}&=&-\frac{\beta}{2\gamma}\pm\sqrt{\left(\frac{\beta}{\gamma}\right)^2-\frac{4(2D_0-C_0)}{\gamma}}\nonumber
\label{dof_phi_ss_soln}
\end{eqnarray}
At first sight this exponential solution might seem to be a cause for concern, as we would normally expect the growth of $\Phi$ or $\delta_M$ on small scales to follow power-law behaviour. The unusual solution above has arisen because the modifications in eqns.(\ref{2nd_order_Us}) introduce factors of $k$ that dominate over the usual GR terms on small scales. In the toy model considered here we took the MGFs to be dimensionless functions of order one; without such assumptions it is difficult to make any general statements about the growth of perturbations, because we do not know how to assess the relative importance of two factors such as $J_0$ and $3k/\ensuremath{{\cal H}}$. Of course, if we know the specific functional forms hidden behind the MGFs then such assumptions are not necessary, but then we would not be pursuing a model-independent approach.
A more realistic situation would be to take the MGFs to be much smaller in magnitude than the GR terms. However, our intention here is to assess qualitatively the effects that parameterized systems of modifications to gravity have on the growth of structure - a task which becomes difficult when the parameters are taken to be vanishingly small. We will therefore maintain the simple assumptions described above, remembering that in a more realistic model the effects described here would only be manifest as small distortions of a predominantly GR-controlled universe.
Alternatively, we could turn this problem on its head. One reason theories with additional d.o.f. are difficult to work with is because we lose the ability to derive a hierarchy of constraint equations between the MGFs, as we did in the purely metric case (these arise from the Bianchi identities -- see appendix \ref{app:Z_consv}). Can we use the growth of $\Phi(x)$ to infer some replacement constraints?
To clarify, we wish to restore power-law behaviour for $\Phi(x)$. By expanding eqn.(\ref{dof_phi_eqn}) in powers of $k/\ensuremath{{\cal H}}$ we can find the conditions necessary to remove the dominant terms that are causing the exponential solution of eqn.(\ref{dof_phi_ss_soln}). We find these to be:
\begin{eqnarray}
2D_0-C_0=0,\;\quad2D_1-C_1=0,\;\quad2K_0-J_0=0
\label{constraints}
\end{eqnarray}
If the above conditions are satisfied on small scales then eqn.(\ref{dof_phi_eqn}) reduces to:
\begin{eqnarray}
&&\ddot\Phi+\ensuremath{{\cal H}}\dot\Phi\left[1-D_0+\frac{12}{\gamma}\right]+\frac{\ensuremath{{\cal H}}^2}{2}\Phi \left(1-D_0\right)\left[\frac{6}{\gamma}-1\right]=0\nonumber\\
\end{eqnarray}
The solutions are then power laws in $a$ (or $\eta$), as desired:
\begin{eqnarray}
\label{phi_sol_dof_sub}
\Phi (a)&=&Pa^{\frac{q^+}{2}}+Qa^{\frac{q^-}{2}}\nonumber\\
q^{\pm}&=&D_0-\frac{12}{\gamma}-\frac{1}{2}\nonumber\\
&&\pm\left(D_0^2-3D_0-\frac{12 D_0}{\gamma}+\frac{144}{\gamma^2}+\frac{9}{4}\right)^{\frac{1}{2}}
\end{eqnarray}
In appendix \ref{app:deltas_in_dof} we convert this solution for $\Phi (a)$ into a solution for $\delta_M$ and $\delta_E$. Here we shall simply state the results:
\begin{eqnarray}
\delta_M (a)&=&P k^2\frac{(D_0-1) }{(q^{+}+2)(q^{+}+3)}a^{\frac{q^{+}}{2}+1}\\
\delta_E (a)&=&\frac{P k^2}{6\Omega_E}\left[\frac{6(1-D_0)(1-\Omega_E)}{(q^{+}+2)(q^{+}+3)}-\frac{1}{\mu_P}\right] a^{\frac{q^{+}}{2}+1}
\end{eqnarray}
In the GR limit $\Omega_E\to 0$, $D_0\to 0$ (which sends $q^{+}\to 0$) we recover that $\delta_M$ scales with $a$.
It is interesting to see that the conditions in eqn.(\ref{constraints}) are satisfied in parameterization A of \textsection\ref{section:param_choice} in the limit $k\to\infty$ (the last condition trivially so). The same is true in parameterization B, where $D_1=0$.
If one wished to implement this gravitational framework into a realistic cosmological model, without necessarily taking the absolute magnitude of the MGFs to be very small, then these are the constraints that must be satisfied to give a reasonable degree of structure formation. Of course, this does not prevent the model from being ruled out by other observables such as the ISW effect. Having restored power-law growth, the effects of this theory on the ISW effect and growth function are expected to be qualitatively similar to those in \textsection\ref{sub:growth_observables}.
\subsection{Superhorizon scales}
\label{subsub:dof_superhorizon}
On superhorizon scales eqn.(\ref{dof_phi_eqn}) simplifies to:
\begin{eqnarray}
&&\ddot\Phi+\frac{k\dot\Phi}{K_1}\left(K_0-\frac{D_1}{3}\right)\\
&&-\frac{k^2\Phi}{18 K_1}\Big[(J_1-12)(D_0-1)+K_1\left(2-C_0\right)\Big]=0\nonumber
\label{dof_large_scales}
\end{eqnarray}
This can be reduced to an oscillator equation in $k\eta$, in complete analogy to the subhorizon case. However, in the superhorizon limit $k\to\infty$ these oscillations become infinitely slow in $k\eta$; effectively, we have that $\Phi$ is a constant. This matches the GR and metric-only cases.
\section{Conserved Superhorizon Perturbations?}
\label{section:zeta}
Any relativistic theory of gravity in which energy-momentum is covariantly conserved allows the definition of a perturbation, ${\cal Z}$, that is conserved on superhorizon scales in the absence of non-adiabatic perturbations \cite{Wands_etal_2000, Cardoso_Wands, Bertschinger2006}. In the literature ${\cal Z}$ is more commonly denoted as $\zeta$, but the choice of notation for one of the MGFs in this paper and previous work prevents us from reusing that letter. ${\cal Z}$ is identified with the curvature perturbation on uniform-expansion hypersurfaces in a homogeneous and isotropic spacetime, which in GR coincide with hypersurfaces of constant energy density. For zeroth-order Einstein equations of the form:
\begin{eqnarray}
{\cal H}^2&=&\frac{a^2}{3}f_0\nonumber\\
\dot{\cal H}-{\cal H}^2&=&-\frac{a^2}{2}g_0
\label{zero_order_fg}
\end{eqnarray}
the conserved perturbation is \cite{Cardoso_Wands}:
\begin{equation}
\label{zeta_definition}
{\cal Z}=-\Phi-\frac{\cal H}{\dot f_0}\delta f
\end{equation}
It would be interesting to know how ${\cal Z}$ behaves in the metric theories considered in \textsection\ref{section:delta}.
Since the new non-GR terms do not really originate from perturbations to a fluid, it is not immediately obvious whether they will be equivalent to adiabatic or non-adiabatic pressure perturbations. Therefore the conservation of ${\cal Z}$ does not necessarily follow.\newline
Using the linearly perturbed versions of eqns.({\ref{zero_order_fg}}), one can derive an equation for the evolution of the metric potentials \cite{Cardoso_Wands}:
\begin{equation}
\label{ode}
\ddot\Phi+\frac{3{\cal H}\dot{\cal H}-\ddot{\cal H}-{\cal H}^3}{\dot{\cal H}-{\cal H}^2}\dot\Phi+{\cal H}\dot\Psi+\frac{2\dot{\cal H}^2-{\cal H}\ddot{\cal H}}{\dot{\cal H}-{\cal H}^2}\Psi=\frac{a^2}{2}\delta g_{nad}
\end{equation}
where the perturbation $\delta g$ has been decomposed into parts equivalent to adiabatic and non-adiabatic pressure perturbations:
\begin{equation}
\label{delta_g}
\delta g=\frac{\dot g_0}{\dot f_0}\delta f+\delta g_{nad}
\end{equation}
The time derivative of ${\cal Z}$ is related to the quantity on the LHS of eqn.(\ref{ode}). Hence the rate of change of ${\cal Z}$ is found to be:
\begin{equation}
\label{zeta_dot}
\dot{\cal Z}=\frac{\cal H}{\dot{\cal H}-{\cal H}^2}\frac{a^2}{2}\delta g_{nad}
\end{equation}
Comparing eqns.(\ref{einstein}) and (\ref{zero_order_fg}) and defining \mbox{$U_{00}=X$}, \mbox{$U_{ii}=Y$} we can read off:
\begin{eqnarray}
f_0&=&\kappa \rho_M+X\label{f_0}\nonumber\\
g_0&=&\kappa (\rho_M+P_M)+X+Y\label{g_0}
\end{eqnarray}
Rearranging eq.(\ref{delta_g}) and substituting in our expressions for $f$ and $g$, one finds:
\begin{eqnarray}
\hspace{-0.3cm}\delta g_{nad}&=&\frac{\kappa\dot\rho_M\dot X}{\kappa\dot\rho_M+\dot X}\left(c_s^{2(M)}-c_s^{2(X)}\right)\Gamma_{\rho_M X}+\kappa \delta P_{nad}+\delta Y_{nad}\nonumber\\
\label{delta_gnad}
\end{eqnarray}
where
\begin{eqnarray}
\Gamma_{\rho_M X}&=&\frac{\delta\rho_M}{\dot\rho_M}-\frac{\delta X}{\dot X}\nonumber\\
c_s^{2(M)}&=&\frac{\dot P_M}{\dot\rho_M}\nonumber\\
c_s^{2(X)}&=&\frac{\dot Y}{\dot X}\nonumber
\end{eqnarray}
The perturbations $\delta P_M$ and $\delta Y$ have been decomposed in a manner analogous to eq.(\ref{delta_g}). $\Gamma_{\rho_M X}$ represents a possible entropy perturbation between CDM and the modified sector, which can be non-zero even if each component does not support entropy perturbations by itself.
Perturbations of the background quantities $X$ and $Y$ correspond to components of the tensor $U_{\mu\nu}$: \mbox{$a^2 \delta X=-a^2 \delta U_0^0$} and \mbox{$a^2\delta Y=a^2\delta U^i_i/3$}. Then the effective non-adiabatic pressure perturbation of the modified sector is:
\begin{equation}
\label{Y_nad}
\delta Y_{nad}=\delta Y-c_s^{2(X)}\delta X=\frac{\delta U_i^i}{3}+c_s^{2(X)}\delta U_0^0
\end{equation}
Our toy model of an EdS universe contains only cold dark matter, so \mbox{$\delta P_M=\delta P_{nad}=c_s^{2(M)}=0$}.
\subsection{Metric-Only Theories}
For a purely metric theory with a $\Lambda$CDM-like background $X$ and $Y$ are zero or equivalent to a cosmological constant (i.e. \mbox{$X+Y=0$}), so there can be no entropy perturbations \textit{between} CDM and the modifications (since \mbox{$c_s^{2(X)}=0$}). The only possible source for non-conservation of $\cal Z$ would be from non-adiabatic perturbations within the modified sector itself, $\delta Y_{\mathrm{nad}}$. Eqn.(\ref{zeta_dot}) becomes:
\begin{eqnarray}
\label{zeta_dot_metric}
\dot{\cal Z}&=&\frac{\cal H}{2(\dot{\cal H}-{\cal H}^2)}\frac{a^2\delta U_i^i}{3}\nonumber\\
&=&\frac{\cal H}{6(\dot{\cal H}-{\cal H}^2)}\left(k^2C_0\Phi+kC_1\dot\Phi\right)
\end{eqnarray}
Non-adiabatic perturbations within the modified sector would amount to fluctuations about $\omega_E=-1$, which could lead to a situation equivalent to a phantom field. Whilst a phantom equation of state is permitted by current data \cite{WMAP7_Komatsu, Nesseris_Perivo,Perivolaropoulos}, direct phantom scalar field models are plagued by severe difficulties because they lead to an unstable vacuum state \cite{Cline_2004} and favour an anisotropic universe \cite{Gannouji_Polarski}. However, it is known that theories such as scalar-tensor gravity, $f(R)$ gravity and some Lorentz-violating models can cause phases of an effective $\omega_E<-1$ without introducing a phantom field \textit{per se} \cite{Abdalla_Nojiri_Odintsov, Amendola_Tsujikawa, Martin_Schimd_Uzan, Libanov_Rubakov, Kunz_Sapone_phantom}.
Fortunately the question of whether non-adiabatic perturbations lead to \mbox{$\omega_E<-1$} turns out to be a moot point for the theories considered in \textsection\ref{section:param_choice} and \textsection\ref{section:delta}. This is because the effective pressure perturbation in eqn.(\ref{Y_nad}) vanishes on very large scales. The functions $C_0$ and $C_1$ are related to $\mu_P$ and $\zeta$ by a set of constraint equations, displayed in appendix \ref{app:Z_consv}. There we also show that under the assumptions mades in this paper \mbox{$\lim_{k\to 0}\left(k^2 C_0\right)=0$} and \mbox{$\lim_{k\to 0}\left(k C_1\right)=0$} in both of the parameterizations described in \textsection\ref{section:param_choice} . These results agree with the conclusions of \cite{Pogosian_Silvestri}.
\subsection{Theories with Additional Degrees of Freedom}
For theories with extra scalar degrees of freedom the situation is different. Firstly, the final term in $\delta Y_{nad}$ does not vanish, as the modifications to the background equations lead to $c_s^{2(X)}\neq 0$. Secondly, we cannot derive a system of constraint equations on the functions $A_0\ldots K_1$ like those in Table \ref{tab:2nd_order_constraints}, because the new scalar now acts as a source in the Bianchi identities (see appendix \ref{app:Z_consv} for details). Thirdly, we can now have entropy perturbations between matter and the modified sector, as the coefficient of $\Gamma_{\rho_M X}$ in eqn.(\ref{delta_gnad}) no longer vanishes.
Maintaining the approach of \textsection\ref{section:extra_dof}, we keep the metric components of $\delta U_{\mu\nu}$ distinct, but treat the perturbations of the new d.o.f. as an effective fluid. We previously set the equation of state of this effective fluid to zero in order to preserve the EdS nature of our toy model; let us restore the general case for the present. The full expression for $\dot{\cal Z}$ then becomes:
\begin{widetext}
\begin{eqnarray}
\label{zeta_dot_general}
\dot{\cal Z}&=&\frac{\cal H}{2\left(\dot{\cal H}-{\cal H}^2\right)}\Bigg[\frac{\kappa \dot{\rho_M}\dot{X}}{\kappa\dot{\rho}_M+\dot{X}}\,c_s^{2(X)}\left(\frac{a^2\delta_M}{3\ensuremath{{\cal H}}}+\frac{\kappa a^2 \rho_E\delta_E}{\dot X}\right)+\kappa a^2 \rho_E\left(\Pi_E-c_s^{2(X)}\delta_E\right)\nonumber\\
&&+k^2\Phi\left(\frac{1}{3}C_0-A_0\,c_s^{2(X)}\frac{\dot X}{\kappa\dot{\rho}_M+\dot{X}}\right)+k^2\Gamma\left(\frac{1}{3}J_0-F_0\, c_s^{2(X)}\frac{\dot X}{\kappa\dot{\rho}_M+\dot{X}}\right)+\frac{1}{3}k\left(C_1\dot\Phi+J_1\dot\Gamma\right)\Bigg]
\end{eqnarray}
\end{widetext}
If the MGFs do not contain inverse powers of $k$ then the second line vanishes on very large scales, leaving only perturbations to the effective fluid (although a similar assumption was used in \textsection\ref{section:extra_dof} to make an analytic solution achievable, it does not necessarily have to hold true for all theories).
The first set of round brackets represents entropy perturbations between CDM and the modified sector, and the second term is akin to non-adiabatic perturbations within the modified sector itself. However, note that $c_s^{2(X)}$ is not necessarily equal to the sound speed of the effective fluid, since $X$ and $Y$ may contain terms constructed from the metric. This feature distinguishes a dark energy model like quintessence from a modified gravity model such as a scalar-tensor gravity. In the latter the scalar field is non-trivially coupled to the metric and hence $X$ and $Y$ contain more terms than just the energy density and pressure of a scalar field. A quintessence-like case can be recovered from the above expression by setting the MGFs to zero and \mbox{$X=\kappa\rho_E$}, \mbox{$Y=\kappa\,P_E$}, which gives:
\begin{eqnarray}
\hspace{-1cm}
\dot{\cal Z}&=&-\frac{\cal H}{\left(\dot{\cal H}-{\cal H}^2\right)}\frac{a^2}{2}\Bigg[\frac{\rho_E(1+\omega_E)\,c_s^{2(E)}\Gamma_{M,E}}{1+(1+\omega_E)\frac{\Omega_E}{\Omega_M}}+\kappa\, \delta P_{E, \mathrm{nad}}\Bigg]\nonumber\\
\end{eqnarray}
where
\begin{equation}
\Gamma_{M,E}=\delta_M-\frac{\delta_E}{1+\omega_E}
\end{equation}
If the quintessence field supports only adiabatic perturbations ($\delta P_{E,\mathrm{nad}}=0$) then $\Gamma_{M,E}$ can be set to zero through choice of adiabatic initial conditions, leaving ${\cal Z}$ conserved.
As a second example, consider the theory treated in \textsection\ref{section:extra_dof} in which the effective fluid was assumed to be pressureless. Eqn.(\ref{zeta_dot_general}) reduces to:
\begin{eqnarray}
\dot{\cal Z}&=&-\frac{\cal H}{\left(\dot{\cal H}-{\cal H}^2\right)}\frac{\kappa a^2}{2}\frac{\dot{X}}{\left(\kappa\dot{\rho}_M+\dot{X}\right)}\,c_s^{2(X)}\left[\rho_M\delta_M+\rho_E\delta_E\right]\nonumber\\
\end{eqnarray}
The square brackets can be set to zero at through a choice of isocurvature initial conditions, so $\dot{\cal Z}$ remains conserved if $\omega_E$ is constant. However, as one wishes the effects of modified gravity to become apparent at late times in the universe an evolving $\omega_E$ would be more desirable, for which $\dot{\cal Z}$ would not be conserved.
\section{Conclusions}
\label{section:discussion}
Awareness of a forthcoming `data deluge' from current and future cosmological experiments has led to an interest in model-independent approaches to constraining modified gravity. By exploiting the generic features of current models, these parameterized systems seek to constrain large regions of theory space simultaneously. Whilst there has been much effort made to constrain these parameterizations with the data \cite{Bean, DanielLinder, Zhao2010}, there has been relatively little investigation into the corresponding theoretical description: how does the parameterized system of perturbation equations evolve? (Although see \cite{Ferreira_Skordis}; general scalar field-type models are treated in \cite{Tsujikawa_2007,deFelice_2010,Brax_Davis}).
We have attempted to answer this question within the simplified setting of an Einstein-de Sitter universe, for two classes of theories: 1) those for which the degrees of freedom are the metric and matter perturbations, and 2) theories which explicitly introduce additional scalar degrees of freedom. In both cases we have found that the curvature perturbation remains constant on the very largest scales (well above our observable horizon).
On subhorizon scales we found that in the metric-only case perturbations grow in familiar power-law fashion, but the exponents are modified from their General Relativistic values. The two modified gravity functions $\mu_P$ (controlling the effective Newton's constant) and $\zeta$ (controlling the dominant component of the gravitational slip) are partially degenerate in their effects, making them difficult to disentangle. The impact of $\zeta$ becomes subdominant if the effective Newton's constant is weakened. Modifications to the evolution of $\Phi$ lead to an induced ISW effect and modification to the growth rate, $f(z)$.
Theories with additional scalar degrees of freedom are considerably harder to study, as the number of undetermined functions is much larger in this case. In \textsection\ref{section:extra_dof} we considered phases during which the new terms in the Einstein equations behave as an effective fluid with a negligible equation of state. We found that the modified terms dominate the evolution equations, leading to damped oscillatory or exponential behaviour. From a study of these equations we have found three relations (eqns.(\ref{constraints})) between the modified terms that, if satisfied, will restore power-law behaviour. This analysis does not apply to all modified gravity theories (when $\omega_E$ is not small), but it is relevant to (for example) EdS regimes of linear Einstein-Aether theory, EBI gravity and Ho\v{r}ava-Lifshitz gravity.
Constraint equations such as this are desirable because the number of free functions required to parameterize common gravitational theories increases rapidly when new degrees of freedom are included. This proliferation could reduce our ability to constrain such parameterized frameworks satisfactorily. One can reduce this freedom somewhat by treating the new scalar degrees of freedom as an effective fluid; indeed this (or an equivalent approach) is necessary if the parameterization is to capture theories which introduce more than two new scalar degrees of freedom. The metric perturbations are also a considerable source of freedom, as displayed in eqns.(\ref{2nd_order_Us}). This splitting of the modifications into metric parts and effective fluid parts is the key to distinguishing between closely related models of dark energy and modified gravity, such as quintessence and scalar-tensor theories.
Under the assumptions made in this paper the metric terms become irrelevant on ultra-large scales. In a metric-only theory this leaves the superhorizon perturbations ${\cal Z}$ (more commonly denoted by by $\zeta$) conserved, but in class 2) theories the possible non-conservation of $\cal Z$ depends on the equation of state of the effective fluid.
The EdS model considered in this paper is useful in obtaining a qualitative understanding of how parameterized gravity might effect the matter-dominated phase of our universe. To obtain more quantitive predictions this must be embedded in a more complex cosmological model, which is likely to be achievable only numerically.\newline
\section*{Acknowledgements}
\noindent We acknowledge useful discussions with E. Bertschinger, P. Ferreira, J. Pearson, C. Skordis, D. Wands and J. Zuntz. This work was supported by the STFC.
|
\section[]{Introduction}
\label{Introduction}
The light profiles of spiral galaxies consist of two principal components: an inner, bulge-dominated
component; and an outer exponentially declining disc with some minor deviations related to spiral arms
\citep{deVaucouleurs:1959, Freeman:1970}. However, since \cite{vanderKruit:1979} we have known that this
`classical' picture fails for most spiral galaxies, particularly at the faint surface brightness $\mu$ of
the outer stellar disc. We now know that most disc profiles are best described by a two slope model (broken
exponential), characterised by an inner and outer exponential scalelength separated by a relatively well
defined break radius $r_{\rm brk}$ \citep{Pohlen_etal:2002}. Many studies have now reported (mainly using
surface photometry) the existence of broken exponential discs, or {\em truncations}, in spiral galaxies in
both the local \citep{Pohlen_etal:2002, Pohlen_etal:2007, Pohlen_Trujillo:2006, Bakos_etal:2008,
Erwin_etal:2008, Gutierrez_etal:2011, Maltby_etal:2011} and distant $z<1$ Universe \citep{Perez:2004,
Trujillo_Pohlen:2005, Azzollini_etal:2008}. Broken exponential discs have also been reported using resolved
star counts on some nearby galaxies \citep{Ibata_etal:2005,Ferguson_etal:2007}.
These studies have resulted in a comprehensive classification scheme for disc galaxies based on break
features in the outer disc component of their radial $\mu$ profiles \citep[see e.g.][]{Erwin_etal:2005,
Erwin_etal:2008, Pohlen_Trujillo:2006}. This scheme consists of three broad profile types (Type I, II and
III): Type I (no break) -- a single exponential disc extending out to several scalelengths
\citep[e.g.][]{BlandHawthorn_etal:2005}; Type II (down-bending break, {\em truncation}) -- a broken
exponential disc with a shallow inner and steeper outer region \citep{vanderKruit:1979,Pohlen_etal:2002};
Type III (up-bending break, {\em antitruncation}) -- a broken exponential disc with a shallower region beyond
the break radius $r_{\rm brk}$ \citep{Erwin_etal:2005}.
In the classical picture (simple bulge and disc), the de Vaucouleurs, $r^{1/4}$ bulge profile dominates in
the centre while the exponential disc dominates at larger radii. However, theoretically the $r^{1/4}$
profile would dominate again at some low surface brightness. \cite{Erwin_etal:2005} suggest that in some
Type~III profiles (up-bending breaks), the excess light beyond the break radius $r_{\rm brk}$ could
be attributed to light from the spheroidal bulge or halo extending beyond the end of the disc. Consequently,
Type~III profiles can be separated into two distinct sub-classes depending on whether the outer profile
$r>r_{\rm brk}$ is dominated by a disc (Type~III-d) or spheroidal component (Type~III-s).
\cite{Erwin_etal:2005} also propose that antitruncations with a smooth gradual transition and outer isophotes
that are progressively rounder than that of the main disc, suggest an inclined disc embedded in a more
spheroidal outer region such as an extended bulge or halo (i.e. Type~III-s). Using this `ellipse' method,
previous works \citep{Erwin_etal:2005, Erwin_etal:2008, Gutierrez_etal:2011} have found that $\sim40$ per
cent of their Type~III profiles are Type~III-s.
However, the ellipse method is limited for face-on discs and cases where the outer/inner disc may have
different orientations and axis ratios. In these instances, bulge--disc \mbox{(B-D)} decomposition
\citep[e.g.][]{Allen_etal:2006} provides a useful tool to determine the contribution of the two major
structural components (bulge and disc) to the galaxy's light distribution and should provide more conclusive
evidence. The aim of this work is to use B-D decomposition on a large sample of $78$ outer disc
antitruncations from the Space Telescope A901/2 Galaxy Evolution Survey \citep[STAGES;][]{Gray_etal:2009} and
assess the fraction of Type~III profiles that show evidence for the excess light at large radii being caused
or affected by the spheroidal component. This work builds on previous studies by using an improved method for
the classification of Type III-s/III-d profiles (especially for face-on discs) and by using a larger more
representative sample spanning the range of spiral morphologies.
Throughout this paper, we adopt a cosmology of $H_0=70\,{\rm kms^{-1}Mpc^{-1}}$, $\Omega_\Lambda=0.7$, and
$\Omega_m=0.3$, and use AB magnitudes unless stated otherwise.
\section[]{Data and Sample Selection}
\label{Data and Sample Selection}
This work is entirely based on the STAGES data published by \cite{Gray_etal:2009}. STAGES is an extensive
multiwavelength survey targeting the Abell 901/2 multicluster system ($z\sim0.167$) and covering a wide
range of galaxy environments. {\em Hubble Space Telescope} ({\em HST})/Advanced Camera for Surveys (ACS)
$V$-band (F606W) imaging covering the full $0.5^{\circ}\times0.5^{\circ}$ ($\sim5\times5\,{\rm Mpc^2}$) span
of the multicluster system is complemented by extensive observations including photometric redshifts from
\cite{Wolf_etal:2003}. All imaging and data are publicly
available\footnote{http://www.nottingham.ac.uk/astronomy/stages}. In addition to this, all galaxies with
apparent $R_{\rm vega} < 23.5\,{\rm mag}$ and $z_{\rm phot} < 0.4$ (5090 galaxies) were visually
classified by seven members of the STAGES team into the Hubble types (E, S0, Sa, Sb, Sc, Sd, Irr)
and their intermediate classes (Gray et al. in prep.).
Our galaxy sample is drawn from \cite{Maltby_etal:2011}. This consists of a large, mass-limited
($M_* > 10^9\,\rm M_\odot$), visually classified (Sa-Sdm, from Gray et al. in prep) sample of $327$
face-on to intermediate inclined ($i< 60\,\rm degrees$) spiral galaxies from both the field and cluster
environments. However, we remove two galaxies for which B-D decomposition fails ($N_{\rm tot} = 325$). The
$182$ cluster spirals are at a redshift of $z_{\rm cl} = 0.167$ and the $143$ field spirals span a redshift
range of $0.05 < z_{\rm phot} < 0.30$. \cite{Maltby_etal:2011} analysed the $\mu(r)$ profiles for these
galaxies in order to identify broken exponentials in the outer stellar disc $\mu > 24\,\rm mag\,arcsec^{-2}$
(their criteria for selecting intrinsically similar outer breaks). Three independent assessors agreed on a
sub-sample of $78$ antitruncated (Type~III) outer $\mu$ profiles ($\mu_{\rm brk} > 24\,\rm mag\,arcsec^{-2}$).
We shall use both this Type~III sub-sample and the total sample in this work.
\section{Methodology}
\label{Methodology}
For each galaxy in our sample, we perform a two-dimensional B-D decomposition based on a two component
galaxy model comprising of a de Vaucouleurs ($r^{1/4}$) bulge and a single exponential disc. Decompositions
were carried out on the STAGES $V$-band imaging using the {\sc galfit} code \citep{Peng_etal:2002} and the
method of \cite{Hoyos_etal:2011} adapted to perform two component fits. Several measurable properties are
produced for each galaxy including: position [$x$,$y$], effective radii, total magnitudes, axis ratios,
position angles for the bulge and disc components, and a sky-level estimation.
B-D decomposition can be sensitive to the initial conditions used to search the B-D parameter space (e.g.
initial guess for bulge-to-disc ratio $B/D$). Therefore, we perform two runs of the B-D decomposition with
different initial conditions from the two extremes. One run starting from a bulge-dominated system
($B/D = 9$), and the other run starting from a disc-dominated system ($B/D = 1/9$). Comparison of these runs
(hereafter, Run~$1$ and Run~$2$) allows for an assessment of the uniqueness/stability of B-D decomposition
on a galaxy-galaxy basis. In the vast majority of cases ($\sim85$ per cent) the results were effectively the
same, $\sim70$ per cent being exactly the same and $\sim15$ per cent having only minor differences that do
not affect our analysis ($B/D$ the same within $\sim10$ per cent). In a few cases ($\sim10$ per cent) the
decomposition was catastrophically unstable with Run 1/2 yielding both bulge- and disc-dominated systems. The
remaining cases ($\sim5$ per cent) showed moderate instabilities great enough to affect the assessment of
bulge light in the outer regions of the galaxy. These stability fractions are the same in our
Type~III sub-sample. The unstable solutions are mainly driven by differences in the sky level determined
during the decomposition. However, the overall conclusions of this work are not affected by these unstable
solutions.
For each galaxy, we also use the {\sc iraf} task {\em ellipse} (STSDAS package - version 2.12.2) in order to
obtain azimuthally averaged radial $\mu(r)$ profiles from the STAGES $V$-band imaging, see
\cite{Maltby_etal:2011} for full details of the {\em ellipse} fitting method used. For all our {\em ellipse}
fits the galaxy centre is fixed (all isophotes have a common centre) using the galaxy centre determined
during B-D decomposition. The bad pixel masks of \cite{Gray_etal:2009} are also used to remove everything not
associated with the galaxy itself from the isophotal fit (e.g. light from companion galaxies).
Using a similar procedure to previous works \citep{Pohlen_Trujillo:2006,Erwin_etal:2008}, we fit two
different sets of ellipses to each galaxy image. The first is a free parameter fit (fixed centre, free
ellipticity $e$ and position angle ${\it PA}$) and tends to follow morphological features such as bars and
spiral arms. Consequently, these free-fits are not suitable for the characterisation of the underlying outer
disc, however they may be used to determine the $e$ and ${\it PA}$ of the outer disc component, see
\cite{Maltby_etal:2011} for further details. A fixed-parameter fit (fixed centre, $e$ and $\it{PA}$ using the
$e$ and $\it PA$ determined for the outer disc) is then used to produce our final measured $\mu(r)$
profiles. The necessary sky subtraction is then performed using the sky-level estimates generated during B-D
decomposition. Please note that these sky-values sometimes differ slightly from those of
\cite{Maltby_etal:2011}.
Analogous fixed-parameter fits (using the $e$ and $\it PA$ determined for the outer disc) are also carried
out on the disc-residual images (ACS image - bulge-only model) resulting in a measured $\mu$ profile for the
disc component $\mu_{\rm Disc}(r)$. We also obtain azimuthally averaged radial $\mu$ profiles for the
decomposed B-D model using the same fixed-parameter ellipses (isophotes) as in the other profiles. These
yield separate radial $\mu$ profiles for both the disc and bulge model along the semi-major axis of the outer
disc.
All resultant $\mu$ profiles are corrected for Galactic foreground extinction, individual galaxy inclination
$i$, and surface brightness dimming (the $\mu$ profiles of field galaxies, $0.05 < z_{\rm phot} < 0.30$, are
corrected to the redshift of the cluster $z = 0.167$). Full details of the fitting procedure, subsequent
photometric calibration, and an estimation of the error in the sky subtraction ($\pm0.18$ counts) can be
found in \cite{Maltby_etal:2011}.
\section[]{Results}
\label{Results}
B-D decompositions using a de Vaucouleurs ($r^{1/4}$) bulge plus an exponential disc can be classified into
$4$ distinct profile types \citep[e.g.][]{Allen_etal:2006}, see Fig.~\ref{B-D profile types}.
\begin{enumerate}
\renewcommand{\theenumi}{(\arabic{enumi})}
\item {\em Type A:} `Classical' system. The bulge profile dominates at the centre, while the disc profile
dominates at large radii. The bulge/disc profiles cross only once.
\item {\em Type B:} Disc-dominated system. The disc profile dominates at all radii, with a weak contribution
from the bulge in the centre. The bulge/disc profiles never cross.
\item {\em Type C:} The bulge profile dominates at small/large radii, but the disc profile dominates at
intermediate radii.
\item {\em Type D:} Bulge-dominated system. The bulge profile dominates at all radii with a weak underlying
disc component. The bulge/disc profiles never cross\footnote{Note the use of capital letters in profile
types to avoid confusion with other classification schemes.}.
\end{enumerate}
In addition to this, we also observe decompositions where the disc profile dominates in the centre, while
the bulge profile dominates at large radii (hereafter Type E). Here, we believe an outer antitruncated disk
has incorrectly affected the bulge profile fit. In these cases, B-D decomposition is not a true
representation of the galaxy at large radii and these galaxies probably have Type B-like compositions.
Similar constraints may also occur in some Type~D profiles.
Fig.~\ref{B-D profile types} shows the distribution of B-D profile types for both the total sample and
the Type III sub-sample. As expected, the fraction of Type~C/E profiles is greater in the Type~III
sub-sample. For the total sample, several other correlations between B-D profile type, Hubble-type
morphology (Sa-Sd) and measured $B/D$ ratio are observed. Type A/B profiles are equally probable in all
Hubble types but Type C/D profiles are more common in earlier Hubble types (Sa-Sb). Also as expected,
mean/median $B/D$ decreases with progressively later Hubble types and increases for the sequence of B-D
profile types B--A--C--D (increasing bulge dominance).
\begin{figure}
\centering
\includegraphics[width=0.2\textwidth]{fig_1_top_left.eps}
\includegraphics[width=0.2\textwidth]{fig_1_top_right.eps}\\
\includegraphics[width=0.2\textwidth]{fig_1_middle_left.eps}
\includegraphics[width=0.2\textwidth]{fig_1_middle_right.eps}\\
\includegraphics[width=0.2\textwidth]{fig_1_bottom_left.eps}
\includegraphics[width=0.2\textwidth]{fig_1_bottom_right.eps}
\caption{\label{B-D profile types} Bulge-disc (B-D) profile types. Top left: Type~A, `classical' profile.
Top right: Type~B, disc-dominated. Middle left: Type~C, bulge-dominated at small/large radii but
disc-dominated at intermediate radii. Middle right: Type~D, bulge-dominated. Bottom left: Type~E,
`constrained' outer bulge. Bottom right: profile type distributions for the total sample and Type~III
sub-sample.}
\end{figure}
For each galaxy in our Type~III sub-sample, we compared the measured, fixed ellipse $\mu(r)$ profile
with the model $\mu$ profiles from the B-D model in order to assess the contribution of bulge light in the
outer regions of the galaxy ($r > r_{\rm brk}$). Fig.~\ref{Examples} shows some examples for each B-D profile
type. These comparisons resulted in three possible scenarios. Bulge light in the outer profile
($r>r_{\rm brk}$) either had:
\begin{enumerate}
\item{\em little or no contribution} ($\sim70$ per cent): For all Type~A/B profiles the bulge contributes
virtually no light at $r > r_{\rm brk}$ and in some Type~C/E profiles the contribution is negligible. This
can be determined by inspection of the measured disc-residual profile $\mu_{\rm Disc}(r)$ and assessing if
the properties of the outer profile/break ($r_{\rm brk}$, $\mu_{\rm brk}$, scalelength) have been affected
with respect to the sky-subtraction error. No Type~D profiles are in this category.
\item{\em minor contribution} ($\sim15$ per cent): Approximately half of these cases are Type~C profiles
where the bulge profile emerges from the end of the disc and contributes some light to the outer regions
of the galaxy. The remaining half are Type~E profiles where the bulge appears to be constrained by an outer
exponential disc. The amount contributed is enough to affect the outer profile causing $\mu_{\rm brk}$ and
the outer scalelength to be different in the disc-residual profile $\mu_{\rm Disc}(r)$. However, the
antitruncation remains present.
\item{\em major contribution} ($\sim15$ per cent): Here, the bulge contributes the majority of the light at
$r > r_{\rm brk}$. Approximately half of these cases are Type~D profiles where the bulge dominates at all
radii. For these cases, the de Vaucouleurs profile is either interpreted as an antitruncation or being
constrained by an outer shallow disc. The remaining half are all Type~C profiles (with one exception which
is Type~E). In these latter cases the antitruncation is not observed in the disc-residual profile
$\mu_{\rm Disc}(r)$, see Fig.~\ref{Good_example} for such an example.
\end{enumerate}
\begin{figure*}
\centering
\includegraphics[width=0.24\textwidth]{fig_2_row1_left.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row1_centreleft.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row1_centreright.eps}
\includegraphics[width=0.29\textwidth]{fig_2_row1_right.eps}\\
\includegraphics[width=0.24\textwidth]{fig_2_row2_left.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row2_centreleft.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row2_centreright.eps}
\includegraphics[width=0.29\textwidth]{fig_2_row2_right.eps}\\
\includegraphics[width=0.24\textwidth]{fig_2_row3_left.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row3_centreleft.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row3_centreright.eps}
\includegraphics[width=0.29\textwidth]{fig_2_row3_right.eps}\\
\includegraphics[width=0.24\textwidth]{fig_2_row4_left.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row4_centreleft.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row4_centreright.eps}
\includegraphics[width=0.29\textwidth]{fig_2_row4_right.eps}\\
\includegraphics[width=0.24\textwidth]{fig_2_row5_left.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row5_centreleft.eps}
\includegraphics[width=0.215\textwidth]{fig_2_row5_centreright.eps}
\includegraphics[width=0.29\textwidth]{fig_2_row5_right.eps}\\
\caption{\label{Examples} Example B-D decompositions and antitruncated $\mu$ profiles. Rows, top to bottom:
decompositions producing Type~A, Type~B, Type~C, Type~D and Type~E profiles. The bulge ({\em red dotted line}),
disc ({\em blue line}), and bulge $+$ disc ({\em black dashed line}) profiles from B-D decomposition are
overplotted on the measured $\mu$ profiles ({\em red circles}). The {\em{i}}/{\em{ii}}/{\em{iii}} notation
after the profile type indicates whether the bulge profile has a negligible, minor or major contribution
outside the break radius $r>r_{\rm brk}$, as in Section~\ref{Results}. Errors in the measured $\mu$ profiles
are for an over- and undersubtraction of the sky by $\pm1\sigma$. The $\mu_{\rm crit}$/$\mu_{\rm lim}$ levels
represent $+1\sigma$/$+3\sigma$ above the sky respectively, see \protect\cite{Maltby_etal:2011} for full
details. The model images show the effective radius isophote for both the bulge ({\em red line}) and disc
({\em blue dashed line}) models.}
\end{figure*}
\begin{figure}
\centering
\includegraphics[width=0.44\textwidth]{fig_3.eps}
\caption{\label{Good_example} A rare/unusual example of a bulge profile causing an antitruncation in a $\mu$
profile, (Type~C example from Fig.~\ref{Examples}, break radius $r_{\rm brk}$). The bulge ({\em red dotted
line}), disc ({\em blue line}), and bulge $+$ disc ({\em black dashed line}) profiles from B-D decomposition
are overplotted on the measured $\mu$ profile ({\em red circles}). The disc-residual $\mu_{\rm Disc}$ profile
(measured $\mu$ profile minus bulge-only model, {\em red crosses}) shows no antitruncation (for errors,
see Fig.~\ref{Examples}).}
\end{figure}
\section[]{Conclusions}
\label{Conclusions}
Our results suggest that for the majority of Type~III profiles ($\sim85$ per cent), the excess light beyond
the break radius $r_{\rm brk}$ is related to an outer shallow disc (Type~III-d). However, it is important to
note that for some of these cases ($\sim15$ per cent), bulge light can affect the properties of the disc
profile (e.g. $\mu_{\rm brk}$, outer scalelength). For the remaining Type~III profiles ($\sim15$ per cent),
the excess light at $r>r_{\rm brk}$ can be attributed to the bulge profile (Type~III-s). However, few of
these latter cases (only 3 galaxies with stable decompositions, $\sim5$ per cent) exhibit profiles where the
bulge profile extends beyond a dominant disc (Type~C) and causes an antitruncation in the $\mu$ profile.
Previous works, \citep{Erwin_etal:2008, Gutierrez_etal:2011} have used the ellipse method to classify their
Type~III-d/III-s profiles. The results of these works are in good agreement with their Type~III profiles
being $\sim60$ per cent Type~III-d and $\sim40$ per cent Type~III-s. These results appear {\em not} to
be consistent with our results. However, consider the case where the contribution of bulge light at
$r>r_{\rm brk}$ is too little to explain the antitruncation but great enough to affect the properties of the
profile. The increased contribution of bulge light to the $\mu(r)$ profile at $r_{\rm brk}$ could lead to a
smoothing of the inflection and a roundening of the outer isophotes. Therefore, using the ellipse method
would likely have resulted in these Type~III-d profiles being classified as Type~III-s. Applying this
reasoning to our results, we obtain fractions of $\sim70$/$30$ per cent for Type~III-d/III-s profiles
respectively. Therefore, the ellipse method potentially could lead to genuine disc breaks being classified
as \mbox{Type~III-s}. Our method offers an improved way to determine whether an antitruncation is disc or
spheroid related.
However, our method does have some obvious drawbacks. In a two-component B-D decomposition, an outer
antitruncated disc could cause the bulge profile to be constrained and lead to an over-estimation of bulge
light in the outer regions of the galaxy. This naturally enhances our fraction of Type~III-s profiles which
therefore represents an upper limit to the fraction of genuine Type~III-s profiles in our Type~III sample.
Therefore, we can conclude that in the vast majority of cases Type~III profiles are indeed a true disc
phenomenon.
Several studies have proposed formation scenarios for Type~III profiles, mainly via satellite accretion or
minor mergers \citep{Penarrubia_etal:2006, Younger_etal:2007}. A discussion of these scenarios is beyond the
scope of this work. However, we do make one important comment. Previous works
\citep[e.g.][]{Pohlen_Trujillo:2006, Maltby_etal:2011} have reported that Type~III profiles are more
frequent in earlier Hubble types. This result is consistent with a minor merger scenario for their origin.
However, the excess Type~III profiles in early-types could easily have been related to a natural increase in
the number of Type III-s profiles in earlier Hubble types. Fortunately, we observe exactly the same relations
using just our genuine Type~III-d profiles and thus our results remain consistent with the minor merger
scenario for the formation of Type~III profiles.
\section[]{Acknowledgements}
We wish to thank the anonymous referee for their very careful reading and detailed comments on the original
version of this paper which have helped us to vastly improve it. We also thank Stephen Bamford for useful
discussions. The support for STAGES was provided by NASA through GO-10395 from STScI operated by AURA under
NAS5-26555. DTM and MEG were supported by STFC. CH acknowledges support from a Spanish MEC postdoctoral
fellowship.
\bibliographystyle{mn2e} |
\section{Introduction}
Coulomb blockade of voltage-biased tunnel junctions in the presence of a dissipative environment is a well-understood
phenomenon of quantum transport (for a review see for instance Ref.~\cite{NazarovIngold}).
It has been shown that if the impedance of the environment $R$ is smaller than the
tunnel junction resistance $R_t$, but larger than the quantum of resistance $R_Q=2\pi \hbar /{\rm e} \approx 25.8 k\Omega$
($\hbar$ being the reduced Planck constant and ${\rm e}$ the electron charge),
the current is suppressed for bias voltage $V$ smaller than the Coulomb gap ${\rm e}/2C$, associated with the
capacitance $C$ of the junction ($C$ is typically in the fF range).
This effect can be visible for temperatures lower than the Coulomb charging energy ${\rm e}^2/2C$
and for $R_t \gg R_Q$, which assures suppression of cotunneling.
In practice realizing a high-impedance environment with a flat frequency response till frequencies
of the order of ${\rm e}^2/2C\hbar$ is a challenging experimental problem,
as discussed in detail in Ref.~\cite{IngoldGrabertDevoret2}.
Most experimental observations of Coulomb blockade phenomena in {\em single} tunnel junctions are actually
done in an intermediate-impedance situation ($R\sim R_Q$), which leads at least to clearly non-linear characteristics of the junction;
see for example Refs.~\cite{Popovic,Pierre}.
Nevertheless, it is feasible to realize impedances of the order of some hundreds of k$\Omega$; for instance in the recent
Ref.~\cite{Hongisto} a resistance $R\sim0.4$M$\Omega$ has been realized.
For these values not only the suppression of the current at low-bias voltage should be visible,
but also an appealing effect predicted in the 1980s \cite{GefenBenJacob,AverinLikharev,Likharev-newresults,Likharev-correlated}:
the single-electron tunneling oscillations (SETOs).
The idea behind this effect can be understood in the simplest way in the limit
$R, V \rightarrow \infty$ with $V/R=I_b$, so that the tunnel junction is current biased.
The current slowly charges the capacitance. When
$V=Q/C$ reaches the threshold ${\rm e}/2C$ one electron
can cross the junction. If $I_b \ll {\rm e}/R_t C$
this will happen just after $V$ has reached the threshold.
The charge on the capacitance after the tunneling event will be $Q\simeq-{\rm e}/2$,
and $Q$ will start to slowly increase again.
A time $\sim{\rm e}/I_b$ is needed before a new electron can cross
the junction and the sequence can start again.
The voltage at the junction will thus be periodically modulated
at a tunable frequency $\sim I_b/{\rm e}$.
Observation of this phenomenon is difficult. The implementation of the required strong-impedance environment
has been realized by different authors using on-chip resistors
\cite{Martinis,Kuzmin-usj,Cleland-charge,Cleland-env,PashkinKuzmin,Joyez,Zheng}.
An alternative approach has also been tried by designing the environment with tunnel-junction arrays
\cite{DelsingLikharev,DelsingClaeson}, exploiting the fact that a large number of arrays
reduces the stochastic nature of the current \cite{LikharevBakhvalov}.
Recently, observation of soliton-like single-electron oscillations with this method has been reported \cite{Bylander}.
A similar phenomenon for superconducting Josephson junctions
has been predicted (Bloch oscillations) \cite{LikharevZorin} and investigated
by many authors \cite{KuzminHaviland,Kuzmin,Corlevi2,Hekking}.
Arrays of dc SQUIDs have also been exploited in this case to build up the proper environment to obtain Coulomb blockade of Cooper pairs
\cite{WatanabeHaviland,WatanabeHaviland2,Corlevi}.
Reports on the observation of Shapiro-step-like structures in microwave-irradiated
junctions constitute the present state of the art for the experimental probe of this effect \cite{Kuzmin,PashkinKuzmin,MaibaumZorin}.
Progress in the detection of high-frequency current fluctuations \cite{AguadoKouw,Glattli,Reulet,Deblock}
can open new possibilities of observation of this phenomenon and
of the crossover region, where the oscillations are not completely established.
At the same time the possibility of generating a periodic and frequency-tunable
electric signal without any oscillating source is an interesting opportunity and could have
applications, for instance, as a motion actuator in nanomechanical systems or a controlled
single-electron source.
It should be mentioned that Coulomb blockade is instead easily observed in {\em double} tunnel junctions even in absence of an environment.
This happens since the second tunnel junction plays the role of the large-impedance
environment suppressing quantum charge fluctuations and preventing
tunneling if the voltage is below a threshold.
Note however that this configuration (without environment) will not give rise to single-electron
oscillations, but to the well known sequential transport regime.
Electrons hop stochastically through the first and then the second
junctions: The advantage of an ohmic environment is that it
can generate (at least in principle) a stable current source.
In this paper we study how accurate the SETOs can be as a function of the
impedance and of the bias conditions.
In order to do this we will study the charge-fluctuation spectrum
at the junction capacitance (or equivalently the current-fluctuation
spectrum through the resistance load),
and in particular the width of the peak at the frequency
$\sim I_b/{\rm e}$.
Notwithstanding the relatively large number of papers on this subject,
some specifically addressing the dynamics of arbitrarily biased mesoscopic tunnel junctions
with an analytical approach \cite{Ueda,UedaYamamoto,UedaHatakenaka},
a consistent calculation of these quantities is not available.
It can be useful in order to evaluate the expected effect in view of measuring current
fluctuations in this kind of device.
In this paper we show that the width of the peak scales as the inverse of the
impedance for large $R/R_t$.
The peak remains observable till values of $R/R_t$ of the order
of 5.
The plan of the paper is the following.
In Sec.~\ref{model} we present the model
describing the transport through the
junction.
In Sec.~\ref{rregimes} we discuss the different
regimes that the system undergoes by varying the
bias voltage and the environment resistance.
In particular we obtain an analytical expression for
the the $I$-$V$ characteristic in the SETOs regime.
In Sec.~\ref{SQQ} we calculate analytically
the charge spectrum. The results are
discussed and compared with numerical
Monte Carlo simulations in Sec.~\ref{MCresults}.
Section \ref{conclusions} gives our conclusions.
\section{The system and the model}
\label{model}
Let us consider a tunnel junction
with tunneling resistance $R_t \gg R_Q$
and associated capacitance $C$.
The circuit is voltage biased (at voltage $V_b$)
at zero temperature ($T=0$) in the presence of a resistor of
resistance $R$ in series with the junction (see \reff{Model:fig}, left side).
This circuit is equivalent to one with a current source $I_b$ and a shunt resistor $R_s$ in parallel to the junction,
provided $I_b=V_b/R$ and $R=R_s$ (see \reff{Model:fig}, right side).
We will thus use the parallel configuration to describe the device
in analogy with the previous literature \cite{AverinLikharev}.
It is clear that the results can be readily converted to the voltage
bias case.
In particular note that the limit $R_s\rightarrow \infty$ (specifically $R_s \gg R_t$)
describes the ideal current source.
\begin{figure}\centering
\includegraphics[width=.4\textwidth]{SJshuntedequiv.eps}
\caption{\label{Model:fig}
Circuit scheme of the system considered: a tunnel junction with capacitance
$C$ and tunneling resistance $R_t$, biased by the constant current $I_b$
and shunted by a resistance $R_s$ in parallel (right), which is equivalent to a
junction biased by a voltage $V_b$ and with a resistance $R$ in series (left)
provided that $V_b=I_bR$ and $R=R_s$.
}
\end{figure}
Since we are interested in studying SETOs we need $R_s \gtrsim R_t \gg R_Q$.
We thus assume from the outset that $R_s \gg R_Q$, which allows to neglect
quantum fluctuations and treat the charge degrees of freedom classically \cite{AverinLikharev}.
We also assume that the environment has a flat frequency response
$Z(\omega)=R_s$ up to frequencies $\hbar \omega \approx {\rm e}^2/2C$.
This hypothesis, though not easy to fulfill in practice, is the
common assumption in the literature about this problem, and allows a
simpler and more transparent approach.
In this regime transport through the junction is described by
the theory currently known as orthodox Coulomb blockade theory.
Specifically the electron-tunneling rate depends on the
voltage at the junction ($V_J$) as follows:
\beq
\Gamma(V_J)= \theta(V_J-{\rm e}/2C) (V_J-{\rm e}/2C)/({\rm e}R_t)
\label{rate}
\,,
\eeq
where $\theta$ is the Heaviside function.
[At finite temperature the function $\theta(V)$ is
substituted by $1/(e^{-{\rm e}V/k_B T}-1)$.]
If $(R_Q\!\ll )\; R_s \!\ll \!R_t$ the standard picture of Coulomb blockade applies to the
degrees of freedom of the environment:
They have the time to relax to thermal equilibrium
between two electron-tunneling events.
The current-voltage characteristic in this case is then given by
$I_J(V_J)=(V_J-{\rm e}/2C)\theta(V_J-{\rm e}/2C)$, exposing a clear Coulomb gap
for the current $I_J$ through the junction.
But if $R_s \sim R_t$ or larger (for moderate values of the
bias voltage) the resistive environment (described for instance by
a large collection of bosonic modes) reaches thermal equilibrium,
but not the charge on the capacitance $Q(t)$, which needs
a time $\tau_s=R_s C$ to relax to its stationary state.
Formula \refe{rate} still holds, but with a time-dependent voltage $V(t)=Q(t)/C$.
The time dependence of the charge is given by
the solution of the differential equation:
\beq
\dot Q = -Q/R_sC + I_b
\label{circuit}
\,,
\eeq
which for an initial condition $Q_0$ at $t=0$ reads
\beq
Q_f(Q_0,t)=(Q_0-I_b \tau_s) e^{-t/\tau_s}+I_b \tau_s
\label{Qfunction}
\,.
\eeq
The stochastic problem is then completely formulated and
in the remainder of the paper we discuss the behavior of the
current and of the charge as a function of the two relevant
dimensionless parameters of the problem: $\rho=R_s/R_t$
and $\kappa=(I_b-\Ith)/\Ith $, with $\Ith={\rm e}/2\tau_s$ the
threshold for the current to start flowing through the
tunnel junction.
\begin{figure}\centering
\includegraphics[width=.4\textwidth]{IVpaperBB.eps}
\caption{\label{IjVj:fig}
(Color online) Average current through the junction versus average voltage for different values of the ratio $\rho=R_s/R_t$.
The curves evolve from the standard Coulomb blockade suppression (rightmost line corresponding to $\rho=0.1$)
to a square root behavior (leftmost line $\rho=500$). The dotted line is the large-$\rho$ limit of Eq.~\eqref{SETOextreme}.
}
\end{figure}
The current through the junction has already been calculated numerically
in the very early literature \cite{Likharev-newresults}.
We obtained the same results by Monte Carlo simulation
and, for convenience, we reproduce the curves in \reff{IjVj:fig}.
The limit of infinite $R_s$, or ideal current source, was discussed
in details in Refs.~\cite{AverinLikharev,Zaikin}.
There it was shown in particular that in this limit
the system is in the SETOs regime with frequency $I_b/{\rm e}$
with an averaged voltage at the junction given by (see \cite{AverinLikharev}):
\beq
\langle V_{\rm J}\rangle=\sqrt{{\pi {\rm e} R_t\langle I_{\rm J}\rangle} \over {2C}
\label{SETOextreme}
\,.
\eeq
In the next section we will discuss the behavior of the system
for the intermediate regimes appearing when $R_s/R_t$ is not
infinite, deriving in particular new analytical expressions for the
SETOs frequency and $I$-$V$ characteristics.
\section{Regimes of current transport}
\label{rregimes}
In this section we study the evolution of the current through the junction as a function
of the current bias for different values of the external resistance.
The most interesting case is when $R_s/R_t$ is very large, we thus plot
in Fig.~\ref{regimes:fig} the current on a logarithmic scale for the extreme
value of $R_s/R_t=5 \times 10^2$.
\reff{regimes:fig} will be used as a ``map'' for the rest of the section,
where we will discuss how the junction evolves through the four different regimes
of transport indicated in the figure with roman numerals from (I) to (IV).
\reff{visualcharges:fig} shows the behavior of $Q(t)$ in the different regimes.
\begin{figure}
\includegraphics[width=.45\textwidth]{IbiasVfull5BISBB.eps}
\caption{
(Color online) Average voltage through the junction versus bias current in logarithmic scale ($\rho=5\times10^2$):
Four different transport regimes can be outlined.
The solid line gives the results of the numerical simulations, while dashed lines represent the analytical
curves given in Sec.~\ref{rregimes}.
}
\label{regimes:fig}
\end{figure}
\begin{figure}
\includegraphics[width=.45\textwidth]{visualchargepaperBB.eps}
\caption{
(Color online) The time behavior of the charge in the different regimes.
Extreme values of the parameters $\rho$ and $\kappa$ have been
chosen to best outline the differences.
}
\label{visualcharges:fig}
\end{figure}
\subsection{Non SETOs regimes}
Let us begin with the region indicated with (I) in
\reff{regimes:fig}.
For $I_b<\Ith$ the whole current
flows through the shunting resistance and the voltage
at the junction
\beq
\langle V_{\rm J}\rangle_I=R_s I_b
\eeq
remains below the threshold of the Coulomb blockade: ${\rm e}/2C$.
Note that this first branch of the $I$-$V$ characteristic
in \reff{IjVj:fig} is flattened on the $I_J=0$ value
and it is thus not visible.
For $I_b>\Ith$ transport through the junction becomes possible.
For $I_b-\Ith \rightarrow 0^+$ one can identify
a Poissonian regime of transport (region (II) in \reff{regimes:fig}),
where the time between tunneling events fluctuates strongly.
This is due to the fact that $\Gamma(V)$ as given by
\refE{rate} vanishes linearly
near the threshold, and for very small $I_b-\Ith$ the charge has
always the time (typically $\tau_s$) to reach the saturation value
$\Qs=I_b \tau_s$.
The typical inverse time between two tunnel events is:
\beq
1/\tau_{\rm eff} = \Gamma(\Qs)=(R_s/R_t) (I_b -I_{\rm th})/{\rm e} \ll 1/\tau_s \, .
\eeq
The last inequality sets also the region of existence of the regime (II)
i.e.~, $0<\kappa \ll 1/\rho$.
The average $V_J$ can be readily evaluated
by averaging the oscillations of the charge
$Q(t)=\Qs-{\rm e}\,e^{-t/\tau_s}$
on the average time between two tunneling events
$\tau_{\rm eff}+\tau_s \approx \tau_{\rm eff}$.
This gives:
\beq
\label{II}
\langle V_{\rm J}\rangle_{II}=R_s \left[ I_b(1-R_s/R_t)+I_{\rm th} R_s/R_t\right]
\eeq
and the curve is shown dashed in \reff{regimes:fig}.
Note that the slope changes sign at $R_s/R_t=1$.
We will
see that for $R_s\ll R_t$ this region joins continuously region (IV)
without the appearance of region (III).
It is thus convenient to discuss now the region (IV)
defined as the limit of large $I_b$.
In this limit the junction has a $I$-$V$ characteristic of
a normal resistor shifted by the Coulomb gap.
The average $V_J$ reads then:
\begin{equation}\label{IV}
\langle V_{\rm J}\rangle_{IV}
=\frac{R_sR_t}{R_s+R_t}\left(I_b+\frac{{\rm e}}{2R_tC}\right)
\;.
\end{equation}
This expression holds for $\langle V_{\rm J} \rangle \gg {\rm e}/2C$,
i.e.\ for $I_b/\Ith \gg (R_t+R_s)/R_s$.
For $R_s \gg R_t$ it is then clear that a large region defined by the condition
\beq
1/\rho \ll \kappa \ll \rho
\label{conditionSETO}
\eeq
exists between region (II) and region (IV).
This is the SETOs region, (III) in \reff{regimes:fig}, which will be discussed below.
On the other side, for $R_s\ll R_t$, one sees that region (II) and region (IV)
overlap at $ \kappa \approx 1$.
Actually it is straightforward to check that
\refE{IV} expanded to first order in $R_s/R_t$ coincides with \refE{II}.
\subsection{SETOs regime}
Let us now discuss the single-electron oscillations regime, defined
as region (III) in \reff{regimes:fig}.
This region is present only if $R_s\gg R_t$ and is characterized
by nearly periodic electron tunneling events,
since the time between two events is dominated by the
deterministic charging time of the capacitance.
This time is typically of the order of $t_\star$, defined as
the time needed to charge the capacitance from $Q=-{\rm e}/2$ to $Q={\rm e}/2$:
\beq
\label{tstaradi}
t_\star/\tau_s=
\ln\left(\frac{I_b +I_{th} }{I_b -I_{th}}\right) =\ln {\left(2+\kappa \over \kappa\right)}
\,.
\eeq
The electrons hop just after the threshold voltage has been reached.
A general statistical theoretical framework is presented in
Ref.~\cite{AverinLikharev} (and recalled in the Appendix),
but its analytical solution is given there only in the ideal current
source limit ($R_s/R_t \rightarrow \infty$).
Actually in the case of the SETOs the approach is simplified and
further progress is possible.
In order to obtain these results and for the calculation of the
correlation function of the next section it is convenient to
introduce a few concepts.
\begin{figure}\centering
\includegraphics[width=.45\textwidth]{modeltimestep4.eps}
\caption{\label{modeltimestep}
(Color online) Details of the tunneling process: $t_n$ is the time at which the charge reaches the blockade region border after the $(n-1)$th tunneling \
event and $\delta\tau_n$ is the time it stays outside the border before the $n$th event.
}
\end{figure}
In \reff{modeltimestep} the typical time dependence
of the charge $Q(t)$ in the SETOs regime is shown.
We can associate a number $n$ to each hopping
event and define $t_n$ and $\tau_n=t_n+\delta \tau_n$ as the
instant of time when $Q(t)={\rm e}/2$ and when the hopping
event takes places, respectively (see \reff{modeltimestep}).
These quantities fluctuate randomly, but a correlation
between $t_n$ and $t_{n-1}$ exists.
Inversion of \refE{Qfunction} gives the time needed to reach the
border of the Coulomb blockade region starting from a
charge $Q_0$:
\begin{equation}
\label{xi}
\Xi(Q_0)=-\tau_s \ln\left(\frac{I_b\tau_s-{\rm e}/2}{I_b\tau_s-Q_0}\right)\,.
\end{equation}
The following relation between successive times $t_n$ holds:
\begin{equation}\label{Fgothdef}
t_n-t_{n-1}\equiv\mathfrak{F}(\delta\tau_{n-1})\,,
\end{equation}
with
\begin{eqnarray}
\label{Fgoth}
\mathfrak{F}(\delta\tau)
&=& \delta\tau+\Xi\left(Q_f\left(\frac{{\rm e}}{2},
\delta\tau\right)-{\rm e}\right) \nonumber \\
&=&
\delta \tau-\tau_s \ln\left({\kappa \over 2 + \kappa e^{-\delta \tau/\tau_s}}\right)
\,.
\end{eqnarray}
Let us now introduce the probability $P_n(t)$ that $n$ electrons have tunneled
through the junction at the time $t$.
Within the SETOs region, this quantity is different from zero
only in a small time region of the order of $t_\star$, and, in particular,
the above mentioned condition on the typical hopping time
implies that $P_n(t)$ will reach $1$ and then
vanish in a time much shorter than $t_\star$, just after $Q(t)$ crosses the threshold ${\rm e}/2$.
The rate equation for $t\geq t_n$ [$Q(t_n)={\rm e}/2$] takes the simple form:
\beq
{d P_n\over dt}(t) = -\Gamma(Q_f({\rm e}/2,t)/C) P_n(t)
\eeq
with the initial condition $P_n(t_n)=1$.
The solution reads:
\beq
P_n(t)=e^
-\frac{(I_b \tau_s-{\rm e}/2)}{{\rm e} R_t C}\tau_s \left(\frac{t-t_n}{\tau_s}+
\emph{e}^{-{t-t_n \over \tau_s}}-1\right) }
\label{rateEq}
\,.
\eeq
For short times ($ t-t_n \ll \tau_s$) it has a Gaussian form
\beq
P_n(t) \approx \exp\left\{ -(t-t_n)^2\kappa \rho/(4 \tau_s^2) \right\}
\label{sigmaGaussian}
\eeq
with a decay time scale
$
\sim \tau_s /\sqrt{\kappa\rho} \ll t_\star
$
in region (III).
The Gaussian form will thus be used in the following for the analytical
calculations.
From the knowledge of $P_n(t)$ it is possible to
obtain the probability density that a hopping event takes places
at time $t$: ${\cal P}(t) = - d P_n/dt$, for $t\geq t_n$.
This allows us to calculate the average delay time $\qav{\delta \tau }$ for an electron
to hop after the threshold ${\rm e}/2$ has been crossed by the charge $Q(t)$:
\beq
\qav{\delta \tau } = \int_0^{\infty}\!\!\!\!\!\! dt\; t\; {\cal P}(t)
\,.
\eeq
In particular when $P_n$ can be approximated by
the Gaussian \refe{sigmaGaussian} one obtains
$
\qav{\delta \tau }/\tau_s= \sqrt{\pi/(\kappa \rho)}
$
with $\qav{\delta \tau } \ll t_\star$.
To obtain the period of the SETOs one has to
average the nonlinear expression \refe{Fgoth}:
$
\mathcal{T}=\qav{\Fg}
$, which again in the SETOs region simplifies to
\beq
\mathcal{T}=t_\star + 2\,\qav{\delta \tau }/(2+\kappa) \,.
\label{SETOperiod}
\eeq
Let us now come back to the probability.
Conservation of the probability gives that
$P_{n+1}(t)=1-P_n(t)$.
Since in this approximation the behavior is quasiperiodic,
the charge on the capacitor for the $n+1$ electron is
{\em on average} $Q_f({\rm e}/2,t-t_n-\mathcal{T})$.
(A more precise discussion on the validity of this last average can
be found in the Appendix, where the problem is analyzed more
rigorously.)
The average charge can then be computed
by averaging over a period as follows:
\begin{align}
\langle Q \rangle = \int_{t_n}^{t_n+\mathcal{T}}\!\! \frac{dt}{\mathcal{T}} \; \bigg[& Q_f\left({{\rm e}\over2},t-t_n\right) P_n(t) + \\ \nonumber
+\;& Q_f\left({{\rm e}\over2},t-t_n-\mathcal{T}\right) P_{n+1}(t)\bigg] \,.
\end{align}
In the limit of $\rho \gg 1$ only the gaussian part of the probability
is relevant, and the integral gives:
\begin{align}\label{fullQ}
\langle Q \rangle =&
\Qs -\left(\Qs-{{\rm e}\over2}\right)\left(e^{\mathcal{T} \over \tau_s}-1\right){\tau_s\over \mathcal{T}}\; \times \\ \nonumber
&\times\,\left[
1-\sqrt{ \pi \over \kappa \rho}\;
e^{1 \over \kappa \rho} \; {\rm Erfc}\left({1/\sqrt{\kappa \rho}}\right)
\right]
\,.
\end{align}
\refE{fullQ} leads to $\langle V_J \rangle_{III}=C \langle Q \rangle$.
With little loss in the accuracy \refE{fullQ} can be simplified to the form:
\beq
\langle Q\rangle \approx {{\rm e}\over 2}
\left[\kappa +1 -2 \left(1-\sqrt{ {\pi \over \kappa \rho}}\right) / \ln {\left(2+ \kappa \over \kappa\right)}\right]
\,.
\label{simpleQ}
\eeq
The analytical expressions \refe{fullQ}, \refe{simpleQ}, and \refe{AppQ} obtained
in the appendix, are compared to the Monte Carlo results in
\reff{figNose}.
\begin{figure}\centering
\includegraphics[width=.45\textwidth]{IVnoseBB.eps}
\caption{\label{figNose}
(Color online) Comparison between the different approximations for the
SETOs ``nose'' of the current-voltage characteristics: Monte Carlo data (circles) are shown against
analytical calculations from \refE{AppQ} (solid), \refE{fullQ} (dashed), and \refE{simpleQ} (dot-dashed).
}
\end{figure}
These expressions describe the current with good accuracy, and in particular they
all capture the presence of a minimum in the voltage $V_J$.
This minimum signals the crossover region between two different
kinds of SETOs.
We indicate them in \reff{regimes:fig} as III.1 and III.2.
The latter appears for $1 \ll \kappa \ll \rho$.
In this case, and in the extreme limit $\rho\rightarrow \infty$ the
SETOs period becomes $\mathcal{T}={\rm e}/I_b$, i.e.\ corresponds exactly
to the time needed to the ideal current source to furnish a charge ${\rm e}$.
The saturation value for the charge ($Q_s$) in this regime
is much larger than ${\rm e}/2$, implying that only the linear
part of the exponential in \refE{Qfunction} is explored.
This is important since the small fluctuations in the
hopping times do not affect the evolution equation
for the next electron.
One can readily verify that in the limit of an
ideal current source the charge time dependence
around each $t_n$ is
$Q(t)=I_b (t- n {\rm e}/I_b)$.
The non-linear corrections instead add a
stochastic dependence on the time evolution
of the charge.
The period, for instance, does not depend
on $\qav{\delta \tau}$ anymore for $\kappa \rightarrow \infty$,
as is clear from \refE{SETOperiod}.
This extreme limit is not realistic and thus the second term
given in \refe{SETOperiod} is normally important.
\begin{figure}\centering
\includegraphics[width=.45\textwidth]{regsch2BB.eps}
\caption{
\label{regimescheme:fig}
(Color online) Scheme of the boundaries of the transport regimes as a function of the relevant
parameters: The tunneling/shunt resistances ratio $1/\rho=R_t/R_s$ and the
relative distance of the current bias from the threshold $\kappa=(I_{b}-I_{\rm th})/I_{\rm th}$.
The SETOs regime exists only in the limit $\rho \gg 1$.
}
\end{figure}
Reducing the current bias, the saturation charge $Q_s$ becomes of the order of
${\rm e}/2$ ($\kappa=2Q_s/{\rm e}-1 < 1$) and the non-linear behavior of $Q(t)$
begins to correlate different hopping events.
The charge time evolution in this regime is characteristic and
resembles a shark fin, as shown in \reff{visualcharges:fig}.
It is also clear from \refe{SETOperiod}, that in this regime the stochastic
fluctuations have the greatest
impact on the average SETOs period.
These two regimes can be identified on the current plot \reff{figNose}
and in the analytical expression \refe{simpleQ} as the two branches
joined by a minimum of the voltage.
The large bias behavior ($\kappa\gg 1$) of $\refe{simpleQ}$ gives ${\rm e}\,\sqrt{\pi\kappa/\rho}\,/(2C)$,
which is the Averin-Likharev expression for the current \refe{SETOextreme} ($\langle I_J\rangle \sim I_b$ in this limit),
while in the opposite limit the
long exponential charging time is dominating:
$\qav{V_J}\approx ({\rm e}/C)\sqrt{\pi}/(\sqrt{\kappa \rho} \ln{(2/\kappa)})$.
The overall situation is summarized in \reff{regimescheme:fig}.
Since it is very difficult in practice to experimentally reach large values
of $\rho$, the plot in \reff{regimescheme:fig} suggests
that a good experimental choice can be $\kappa=1$,
for which the SETOs appear for the lowest values of
$\rho$.
We will discuss in the following the correlation function of the charge
in order to analyze quantitatively the evolution of the accuracy of the SETOs.
\section{Charge-fluctuation spectrum}
\label{SQQ}
A quantitative measure of the accuracy of the SETOs is given by
the time correlation of the charge.
This can be defined as
\beq
S(\tau) = \qav{Q(t+\tau) Q(t)} -\qav{Q(t+\tau)}\qav{Q(t)}
\,,
\label{Sdef}
\eeq
where the average is performed over a statistical ensemble
and the result does not depend on $t$, since the stochastic
process is stationary.
Note also that this quantity is proportional to the spectrum
of current fluctuations through the shunt resistance $R_s$:
$S_{R}(\tau)=S(\tau)/\tau_s^2$.
In the case of voltage-biased junction (see \reff{Model:fig}),
$S_{R}$ gives the current fluctuations that can be directly
measured through the load resistance $R$.
For perfectly periodic charge oscillations
the Fourier transform of \refe{Sdef} is given by
a sum of Dirac delta functions at $\omega=2\pi n /\mathcal{T}$,
with $n$ integer.
The non-periodic fluctuations introduce a finite width of these peaks.
The form of $S(\omega)$ measures thus directly and with a simple
procedure the accuracy of the periodic charge transfer.
In this section we derive an analytical expression for $S(\omega)$
that allows us to better understand the origin of the fluctuations.
In the next section we will compare these results to
those obtained numerically by Monte Carlo simulations.
In order to calculate the Fourier transform of $S(\tau)$ for
this stationary process it is
convenient to define the charge $Q(t)$ over a time $0<t<\Lambda$,
with $\Lambda \gg \mathcal{T}$ so that many SETOs are present in a single
sample of $Q(t)$.
One can then calculate the Fourier series
\beq
Q_p = \int_0^{\Lambda} {dt\over \Lambda} Q(t) e^{i p {2\pi\over \Lambda} t }
\quad , \quad
Q(t)=\sum_p e^{-i p {2\pi \over\Lambda} t } Q_p
\label{FourierSeries}
\,.
\eeq
Substituting \refE{FourierSeries} into \refE{Sdef} and averaging over
$t$ one obtains:
\beq
S(\tau)=\sum_{p} \qav{|Q_p|^2} e^{ - i p {2 \pi \over \Lambda} \tau}
-Q_0^2
\,,
\eeq
which can be used numerically to compute the correlation function from the
Monte Carlo data, or analytically, by performing the limit $\Lambda \rightarrow \infty$.
In particular the Fourier transform can be defined as
\beq
S(\omega) = \int_{-\Lambda/2}^{+\Lambda/2} d\tau e^{i\omega \tau -0^+ |\tau|} S(\tau)
\,,
\eeq
which gives
\beq
\label{Somega}
S(\omega) = \sum_{p} (\qav{|Q_p|^2}-\delta_{p,0}Q_0^2) 2 \pi \delta(\omega- \omega_p)
\,,
\eeq
with $\omega_p = 2 \pi p /\Lambda$.
The presence of the Dirac delta functions is an artifact due to the periodic
extension induced by the Fourier transform.
In practice, since the frequency scale $1/\Lambda$ is infinitesimal one can
obtain the smooth function $S(\omega)$ by averaging the expression
\refe{Somega} for each value of $\omega$ over a small interval $2 \pi/\Lambda$.
This simply gives that
\beq\label{Swinkhinc}
S(\omega_p) = \Lambda \qav{|Q_p|^2}
\eeq
for $p\neq 0$ (Wiener-Khinchin theorem).
The problem is now reduced to the calculation of the Fourier series of the charge.
Using the definitions of $t_n$ and $\tau_n$ given before in \refE{xi} and assuming
that the extrema of the time evolution of $Q(t)$ coincide with the two hopping
events at times $\tau_0$ and $\tau_N$ we can write:
\begin{align}
Q_p
=&
\sum_{n=0}^{N-1}
\int_{\tau_n}^{\tau_{n+1}}
{ dt \over \Lambda}
\,Q(t)\,e^{i\omega_p t} =
\\ \nonumber
=&
\sum_{n=0}^{N-1}
{e^{i\omega_p\tau_n}}\!\!
\int_0^{\tau_{n+1}-\tau_n}
\!\!
{ dt \over \Lambda} \;
Q_f({\rm e}/2, t-t_{n+1}+\tau_n) \,e^{i\omega_p t}\,.
\end{align}
In the limit of well-established SETOs the integral gives a contribution
that fluctuates very little. On the contrary the exponentials are much more
sensitive to even small fluctuations of the tunneling times, since the phase
results from the accumulation of many different hopping events.
For this reason we expect that the upper integration limit can
be substituted with the period of the SETOs $\tau_{n+1}-\tau_{n}
\approx \mathcal{T}$ and we use $Q_f({\rm e}/2,t-(\mathcal{T}-\qav{\delta \tau}))$ as
the average charge dependence.
The Fourier transform then takes the form:
\beq
\label{finalSom}
S(\omega) = N \mathcal{T} \qav{|F(\omega)|^2} {\cal A}(\omega)
\,,
\eeq
where
\begin{equation}
\label{Fdef}
F(\omega)
=
{1 \over N} \sum_{n=0}^{N-1} e^{ i \omega t_n}
\end{equation}
and
\beq
\label{Adef}
{\cal A}(\omega) =
\left|e^{i \omega \qav{\delta \tau}} \!\!
\int_0^{\mathcal{T}} \!\!
{dt\over \mathcal{T}} \, Q_f({\rm e}/2,t-\mathcal{T}+\qav{\delta \tau})
e^{i \omega t}
\right|^2
\,.
\eeq
The quantity ${\cal A}$ can be readily evaluated:
\beq\label{Aexplicit}
\begin{split}
{\cal A}&(\omega)= \left({{\rm e}\tau_s\over 2 \mathcal{T} }\right)^2 \times \\
\times & \left|
{\kappa +1 \over \omega \tau_s}(e^{ i\omega \mathcal{T}}-1)
-{\kappa e^{-\qav{\delta \tau}/\tau_s} \over \omega \tau_s +i}(e^{i \omega \mathcal{T}}-e^{ \mathcal{T}/\tau_s})
\right|^2
\,.
\end{split}
\eeq
In order to proceed we have to evaluate also the average
of $F(\omega)$.
It is convenient to express the time at which one event happens
as a sum over the delays between previous events
using \refE{Fgothdef}:
\beq
F(\omega)
={e^{i \omega t_0} \over N}
\left(
1+
\sum_{n=1}^{N-1}
\exp\left\{ i \omega \sum_{k=0}^{n-1} \mathfrak{F}(\delta \tau_k)\right\}
\right) \,.
\eeq
Now the average of $|F(\omega)|^2$ can be performed using the
distribution function ${\cal P}(\delta \tau)$:
\beq
\qav{|F(\omega)|^2}
=
{1\over N}
\left(1+2 {\rm Re} \left\{ {g(\omega) \over 1-g(\omega)}\right\}\right)
+{\delta F\over N^2}
\label{Fform}
\eeq
where we introduce the quantities:
\beq
g(\omega) =\qav{e^{i \omega \mathfrak{F}(\delta \tau)}}
=
\int_0^\infty \!\!\!\!\!\! d (\delta \tau) {\cal P}(\delta \tau) e^{i \omega \mathfrak{F}(\delta \tau)}
\label{gdef}
\eeq
and
$\delta F=2 {\rm Re}\left\{ g(g^{N}-1) /(1-g)^2 \right\}$,
whose contribution to $S(\omega)$ vanishes in the limit $N\rightarrow \infty$.
In conclusion we find:
\beq\label{SomNinfinity}
S(\omega) = \left(1+2 {\rm Re} \left\{ {g(\omega) \over 1-g(\omega)}\right\}\right)
\mathcal{T} {\cal A}(\omega)\,,
\eeq
which constitutes the central result of this section.
We are now in the position to study the spectrum of the charge fluctuations
for the system at hand.
As it can be seen from the form of \refe{Fform} the function has a singularity
for $g\rightarrow 1$.
Since
\beq
|g(\omega)|^2 = \int_0^\infty \!\!\!\!\! dt \int_0^t \!\!\! dt'\; {\cal P}(t) {\cal P}(t')
2 \cos( \omega t) \leq 1
\, ,
\eeq
a singularity is present for $\omega \rightarrow 0$ or
when the fluctuations are negligible so that
$g\approx e^{i \omega \qav{\mathfrak{F}(\delta \tau)}}$.
This picture predicts a series of peaks for the
frequencies $\Omega_n = 2 \pi n/\mathcal{T}$ with $n$ integer.
Small fluctuations introduce a finite width, regularizing
the correlation function.
The numerical integration in the expression of $g(\omega)$ is straightforward,
but it is also possible to obtain an analytical expression.
Deep in the SETOs regime ${\cal P}(\delta \tau)$ has a gaussian
behavior and provides a short cutoff time, so that
we can expand the exponential in \refe{gdef} to second order in
$\delta \tau$.
This gives
\begin{align}\label{gexpand}
g(\omega)& = e^{i \omega t_\star} \; \times \\ \nonumber
\times &\left ( 1 + {2i \omega \over 2+\kappa} \qav{ \delta \tau}
- {\kappa^2 i \omega+4 \omega^2\tau_s
\over 2(2+\kappa)^2 } {\qav{\delta \tau^2}\over \tau_s}
+ \dots \right)
\,.
\end{align}
The maximum of ${\rm Re}\{g/(1-g)\}$ takes place
for ${\rm Arg}[g(\omega)]=0$ that at lowest order in $\qav{\delta \tau}$
gives for the position of the poles:
\beq
\Omega_n
=
{2 \pi n \over t_\star}
\left(1- {2\qav{\delta \tau} \over (2+ \kappa)t_\star }\right)
\,,
\eeq
coinciding at linear order in $\qav{\delta \tau}/\tau_s$
with ${2 \pi n/\mathcal{T}}$.
The phase of $g$ ($g=|g|e^{i \phi}$)
thus vanishes at the minimum of ${\rm Re}\{g/(1-g)\}$, so that near this point
one can write at lowest order $\phi \approx \mathcal{T} (\omega-\Omega_n)$.
The relevant part of \refe{Fform} then reads:
\beq\label{Reg1sug}
{\rm Re}\left\{g \over 1-g\right\}\approx {(1-|g|) \over (1-|g|)^2+ \phi^2 }
\,,
\eeq
and \refE{finalSom} takes the simple Lorentzian form
\beq
S(\omega)\simeq\mathcal{A}(\Omega_n)\,
{\Gamma_n/2 \over \Gamma_n^2/4+(\omega-\Omega_n)^2}
\label{SofOmega}
\,,
\eeq
with the full width at half maximum $\Gamma_n$ defined by:
\beq\label{widthgam}
\Gamma_n=2\;{1-|g| \over \mathcal{T}}={4\Omega_n^2 \over \mathcal{T} (2 + \kappa)^2} \left(\qav{\delta \tau^2}-\qav{\delta \tau}^2\right)\,.
\eeq
The presence in this formula of the mean squared variance of
delay in the tunneling time, $\qav{\;(\delta \tau-\qav{\delta \tau})^2}$,
clearly indicates that the spread in the hopping events controls the width
of the peak, as is physically expected.
One also sees that the width of the poles increases
with $n$.
Performing the average with the gaussian distribution
we find $\qav{\;(\delta \tau-\qav{\delta \tau})^2}/\tau_s^2= 4(1-\pi/4)/(\kappa\rho)$
and for the relative width $\Gamma_n/\Omega_n$ we have the explicit expression:
\beq
\label{DeltaOmega}
{\Gamma_n \over \Omega_n} =
{32 \pi \, n \, (1-\pi/4) \over \rho \, \kappa \, (2+\kappa)^2 \, \ln^2\left[(2+\kappa)/\kappa\right]}
\,.
\eeq
\begin{figure}\centering
\includegraphics[width=.45\textwidth]{spreadkBB.eps}
\caption{\label{spreadk}
The relative width at half height of the first noise peak as a function of $\kappa$, given by \refE{DeltaOmega}, for $\rho=10$.
}
\end{figure}
\refE{DeltaOmega} allows to study the width of the peak in the
charge-fluctuation spectrum, and thus, the accuracy of the SETOs.
It correctly gives that $\Gamma_1 / (2\pi/\mathcal{T}) \ll 1$ in the SETOs regime
($ 1/\rho \ll\kappa \ll \rho$).
It also shows (see \reff{spreadk}) that the \textit{relative} width $\Gamma_1/\Omega_1$ is a
monotonic decreasing function of the bias current ($\kappa$) for a given value of the resistance ($\rho$).
From \refE{SofOmega} it is clear that within this approximation
the weight of the Lorentzian peak is controlled only by the
form factor ${\cal A}(\Omega_n)$.
Using its explicit expression one finds that for large $\rho$ it
becomes independent of $\rho$: ${\cal A}(\Omega_1) \simeq
{\rm e}^2/\left(4\pi^2 + \ln^2\left[\frac{2+\kappa}{\kappa}\right]\right)$.
The full variation takes place in the $\kappa<1$ region, where
the shape of $Q(t)$ evolves from the shark-fin to the sawtooth form.
From the form of \refE{DeltaOmega} one can also see that
for small $\kappa$ the peak broadens and the SETOs are washed
out when $\Gamma_1 \sim \Omega_1$ (for $\kappa \sim 1/\rho$).
For large $\kappa$ the theory instead predicts that the relative
width decreases monotonically \footnote{
This expression does not agree with the expression (57) of
Ref.~\cite{AverinLikharev}, specifically we find a different
functional dependence on $R_s$: $\Gamma \sim 1/R^2_s$ instead
of $1/R_s$.
}:
\beq
\label{DeltaOmega2}
{\Gamma_n \over \Omega_n} =
{8 \pi n (1-\pi/4) \over \rho \kappa }
\,.
\eeq
In this limit the SETOs disappear by a decrease of the weight
of the peak, but within our approximation this is not seen.
Actually for sufficiently large current bias ($\kappa \sim \rho$)
there is a finite probability that a single tunnel event is no more
sufficient to bring the charge back in the Coulomb blockade region.
This is quantified by the value of $P_n(t_n)$, which,
contrary to our hypothesis, can become smaller than $1$.
One can expect the theory to roughly remain valid for
the fraction of tunneling events that leaves $Q(t_n)<{\rm e}/2$.
This describes a peak that remains sharp, but that vanishes
in weight as $P_n(t_n)$.
Another interesting and relevant limit is the low-frequency behavior
of $S(\omega)$.
Expanding \refE{gdef} in $\omega$ one can show that
\begin{align}
N\qav{|F(\omega)|^2}
= \;& {F_2-F_1^2 \over F_1^2}\; + \\ \nonumber
+\; & \omega^2 \; {4F_1 F_2 F_3-F_1^2 F_4 - 3 F_2^3 \over 12 F_1^4}
+ \dots
\end{align}
where $F_n=\qav{\mathfrak{F}^n}$.
If the fluctuations are negligible then $F_n=F_1^n$ and
the noise at low frequency vanishes.
In particular in our case this average takes a simple form if the
explicit expression of $\mathfrak{F}$ is used:
\beq\label{zerofreqfact}
N\qav{|F(\omega)|^2}
\simeq 4 {\qav{\delta\tau^2}-\qav{\delta \tau}^2 \over t_\star^2 (2+\kappa)^2 }
\left(1+{\omega^2 t_\star^2 \over 12}+\dots\right)
\,.
\eeq
We thus find that in the SETOs regime the low-frequency noise is suppressed.
Eq.~\eqref{zerofreqfact} together with the expansion of Eq.~\eqref{Aexplicit} for $\omega\rightarrow 0$
allows us to evaluate the zero-frequency Fano factor:
$\mathbb{F}\equiv {S_J(0)}/{{\rm e}\qav{I_J}}=S(0)/{\tau_s^2{\rm e}\qav{I_J}}$, where $S_J(\omega)$ is the
noise spectrum of the current through the junction and the relation
$S(\omega)=S_J(\omega)\tau_s^2/(1+\omega^2\tau_s^2)$ holds exactly.
The reduction of the current fluctuations in the large-$\rho$ limit
naturally leads to sub-Poissonian noise ($\mathbb{F}<1$)
vanishing for $\rho\rightarrow \infty$:
\begin{equation}
\mathbb{F}
\simeq
\frac{(4 - \pi)(-2 + (\kappa+1)t_\star/\tau_s)^2}
{\rho\,\kappa \,(2 + \kappa)^2 (t_\star/\tau_s)^2}\,.
\end{equation}
Nevertheless this expression is only qualitatively correct,
since the analytical theory
has been designed to describe accurately the noise for frequencies
around the peak of the SETOs.
This will be shown in the next section, where we present numerical simulations.
\section{Numerical simulations}
\label{MCresults}
In this section we show numerical results obtained by Monte Carlo
simulations for the charge-fluctuation spectrum.
The purpose of this section is to compare these results with the
analytical calculations of the preceding sections, valid for $\rho\gg1$ and
$1/\rho \ll \kappa \ll \rho$, and to explore the crossover region
where the SETOs disappear.
The Monte Carlo simulations are performed by generating
different realizations of the stochastic time evolution of the charge $Q(t)$
over a time much longer than the SETOs period.
The time evolution of the charge is obtained by discretizing the time
on a nonuniform grid, such that in the time interval $\Delta t$ the
charge varies by a small quantity and
$P=\Gamma(Q(t)) \Delta t < \mathcal{N} \ll 1 $.
The tunneling event is accepted or refused
by generating a random number between 0 and 1
and by comparing it to $P$.
The deterministic evolution of $Q(t)$ between two events
is simply given by \refE{Qfunction}.
The sequence of time intervals of deterministic evolution interspersed by
tunneling times so constructed gives the full knowledge of $Q(t)$ and
constitutes the stochastic run.
The square modulus of the Fourier transform of the charge is then
easily calculated analytically piecewise, interval by interval.
To obtain the noise as from \refe{Swinkhinc},
just a further average over several runs is needed.
Typically $\mathcal{N}=.01$, each run counts $10^3$ tunneling
events, and an average over $10^4$ realizations is performed.
Let us now discuss the numerical results.
We begin by comparing the form of the first peak in the
noise spectrum.
It is shown scaled by the analytically calculated width
$\Gamma_1$ in \reff{noisewidthcompare} for different
values of $\rho$ and given $\kappa$.
The agreement is excellent.
\begin{figure}
\includegraphics[width=.45\textwidth]{noisewidthcompareBB.eps}
\caption{\label{noisewidthcompare}
(Color online) The noise peak at $\omega=\Omega_1$ obtained from Monte Carlo numerical simulations for values of $\rho$ ranging from $80$ to $200$
in steps of $10$. On the $x$-axis the frequency is shifted by $\Omega_1$ and scaled by the width
$\Gamma_1$ as given by \refE{widthgam}, on the $y$-axis the noise is scaled by $\mathcal{T}\rho$.
}
\end{figure}
We then compare the full $\omega$ dependence of $S(\omega)$ obtained
from \refE{SomNinfinity} with the one calculated numerically.
\begin{figure}
\includegraphics[width=.45\textwidth]{noisecomparepaper2LOGBB.eps}
\caption{\label{noisecompare}
(Color online) Comparison between Monte Carlo data for the charge-fluctuation spectrum and the analytical results obtained from \refE{SomNinfinity}
for different values of $\rho$, at fixed bias condition $\kappa=1$.
}
\end{figure}
We show the comparison in \reff{noisecompare} for the relevant case
$\kappa=1$ (which corresponds to the widest extension in $\rho$ of the SETOs region)
and for $\rho$ ranging from $200$ to $10$.
As expected, the agreement is very good deep in the SETOs regime, for $\rho\gg1$, and at $\rho=10$ all the essential features of the peak
at $\omega=\Omega_1$ are still fairly well represented by the theory.
The analytical approximation correctly finds the main contribution to the
noise, but it tends to underestimate it far from the peak maximum.
\begin{figure}
\includegraphics[width=.45\textwidth]{noise-kvary-rho100BB.eps}\\
\includegraphics[width=.45\textwidth]{noise-kvary-rho10BB.eps}
\caption{\label{noisekvary}
(Color online) Monte Carlo spectra for different bias conditions $\kappa$ at fixed junction environment: $\rho=100$ ($\rho=10$) in the upper (lower) graph.
Increasing $\kappa$ means here moving along a horizontal line in \reff{regimescheme:fig} toward the high-bias boundary of the SETOs region
and it allows us to see how SETOs disappear.
}
\end{figure}
Let us now investigate how the SETOs disappear.
From the experimental point of view a simple parameter that can be varied continuously
is $\kappa\sim I_b$.
We thus plot in \reff{noisekvary} the
evolution of $S(\omega)$ for given $\rho=10$ and 100 as a function of
$\kappa$.
These two plots show several interesting features.
The first striking one is the reduction of the relative widths of the
peaks by increasing $\kappa$.
This is predicted by the analytical expression \eqref{DeltaOmega} and
the numerical calculations assess its validity even outside the region of
applicability of the analytical theory.
Note that in \reff{noisekvary} the frequency axis is rescaled with
$\Omega_1$, thus the apparent weight of the peaks is reduced by
the scaling, but it saturates in the region $\kappa \ll \rho$
as predicted by the analytical theory, and then starts to decrease in the crossover region.
The second visible feature is the appearance of a wide Lorentzian
zero frequency peak that remains the only structure for $\kappa \gg \rho$.
This structure is due to the charge noise induced at the capacitance by the
Poissonian current fluctuations generated by the tunnel junction.
By solving the electromagnetic problem one finds:
$S(\omega) = C^2 |Z(\omega)|^2 e \qav{I_J}$,
where $e \qav{I_J}$ is the standard
tunnel junction Poissonian white noise and
$Z(\omega)=R_\parallel/(1+ i \omega R_\parallel C )$,
with $R_\parallel=R_s R_t/(R_s+R_t)$,
the impedance between the current source and the voltage at the
capacitance.
This gives:
\beq
S(\omega)
=
{{\rm e}\langle I_J\rangle C^2R_s^2 \over 1+2 \rho +
\rho^2(1+ \omega^2R_t^2C^2)}\,,
\eeq
which fits our data very well (not shown).
\begin{figure}
\includegraphics[width=.45\textwidth]{noise-rhovaryBB.eps}
\caption{\label{noiserhovary}
(Color online) Monte Carlo spectra for fixed bias conditions $\kappa=1$ and different values of $\rho$: The crossover from the SETOs regime to
the $\rho\lesssim 1$ region is shown.
}
\end{figure}
Finally we show in \reff{noiserhovary} the evolution of $S(\omega)$
for $\kappa=1$ and for $\rho$ evolving from 1 to 100.
This figure gives an idea of the expected spectrum at the
optimal value $\kappa=1$ as a function of $ \rho$.
It turns out that already at $\rho=3$, $S(\omega)$ presents
a very broad maximum and $\rho=5$ is probably sufficient to observe
a clear structure in $S(\omega)$.
\section{Conclusions}
\label{conclusions}
In this paper we have studied theoretically electronic
transport in a tunnel junction in the presence of a large-resistive environment.
The phenomenon of SETOs has been predicted to appear
in this system for essentially infinite value of the
external resistance, so that the junction can be seen as
current biased.
We investigated under which conditions the SETOs
appear for a realistic finite value of the environment
resistance.
We found analytical expressions for the current
[Eq.~\refe{fullQ}] and for the
charge-fluctuation spectrum [Eq.~\refe{finalSom}
with Eq.~\refe{widthgam}].
Our analytical results describe very well the form of the
peak in the charge noise, which can be regarded as the
hallmark
of the SETOs, since it quantifies the
accuracy of the periodicity in the charge time dependence.
We find that a ratio of $R_s/R_t$ as low as
$5$ can be sufficient to observe a clear structure in
$S(\omega)$ if the bias current is chosen such that
$I_b\approx 2 I_{\rm th} = {\rm e}/R_s C$ (this can be
converted on a condition on the voltage bias
$V_b \approx {\rm e}/C$).
This ratio can be obtained experimentally and thus
in principle SETOs can be observed through the measurement of
the current or charge noise.
The conclusion on the possibility of observing the effect is thus
optimistic.
One should however keep in mind that,
following the literature on this problem, the theory presented
holds at low temperature ($ k_B T \ll {\rm e}^2/2C$) and
neglects quantum fluctuations of the electromagnetic
modes of the environment.
Thermal and quantum fluctuations are expected to have a non-negligible influence
on the transport mechanism at play in the system.
Their evaluation requires a different technical
approach and is beyond the scope of the present paper.
\section*{Acknowledgements}
We are indebted for useful correspondence with Y. Pashkin.
We also gratefully acknowledge fruitful discussions with D. Esteve,
F. Portier, and L.S. Kuzmin.
Comments on the manuscripts are acknowledged from R. Avriller, V. Puller, and M. Houzet.
We finally acknowledge partial financial support through the French
ANR grant QNM No.\ 0404 01.
|
\section{Introduction}\label{sec:intro}
The general aim in quantum tomography is to identify quantum states from measurement outcome statistics.
A collection of observables with this property is called \emph{informationally complete} \cite{Prugovecki77}.
Even a single observable can be informationally complete, but then it must be a noncommutative positive operator valued measure (POVM) \cite{BuLa89,Busch91}.
We will study a class of informationally complete POVMs in dimension $d$ with the minimal number of $d^2$ outcomes and we will explain how they can be implemented as sequential measurements of two $d$-outcome measurements.
One can interpret the generated joint observable as a phase space measurement in the discrete phase space $\Z_d\times\Z_d$ \cite{OpBuBaDr95,Vourdas97}.
There are some particularly interesting approaches to finite dimensional quantum tomography, and one of them is based on complete collections of mutually unbiased bases (MUBs) \cite{Ivanovic81, Wootters86,WoFi89}.
In a $d$-dimensional Hilbert space one needs $d+1$ MUBs in order to be able to identify all quantum states, but it is not known if a complete set of MUBs exists in all dimensions.
In fact, there is evidence that for $d=6$ there is no complete set of MUBs \cite{BrWe08,RaLuEn11}.
In our scheme we start from two mutually unbiased bases connected by the finite Fourier transform.
They define a pair of complementary observables, which cannot be measured jointly.
However, it is possible to realize their joint measurement if some additional noise is allowed.
We show that their joint measurement can be chosen to be informationally complete, and that this can be realized as a sequential measurement where we first perform a `weak measurement' in one basis and then another successive measurement in the other basis.
Compared to the fact that one would need $d+1$ complementary observables in order to reach informational completeness in separate measurements, it is remarkable that in the sequential scheme only two observables suffice.
The price to have a joint measurement is that the marginal observables are not the original complementary observables but their unsharp versions.
We will analyze the required additional noise and characterize the optimal joint observable from this point of view.
The qubit case has been first studied in \cite{Busch86}, and our work generalizes those results to arbitrary finite dimension.
The covariant phase space observables, i.e., POVMs covariant under the finite Weyl-Heisenberg group, play a special role in our investigation. We prove that if a pair of conjugate observables have a joint measurement, then they also have a joint measurement which is a covariant phase space observable.
Since every covariant phase space observable arises from a sequential measurement of two conjugate observables \cite{CaHeTo11}, the covariant phase space observables are an outstanding choice for finite dimensional quantum tomography.
The Weyl-Heisenberg group has also a pivotal role in the investigations of symmetric informationally complete (SIC) observables \cite{ReBlScCa04,Appleby05JMP}.
It is generally believed that a Weyl-Heisenberg covariant SIC observable exists in every finite dimension and their existence is numerically tested in all dimensions up to 67 \cite{ScGr10}.
Our results show that any such observable has a neat sequential realization scheme.
There is an interesting difference between the even and odd dimensional Hilbert spaces.
If we require minimal noise in both marginal observables, then their joint observable is unique.
If $d$ is odd, then this observable is informationally complete.
But if $d$ is even, then the joint observable is not informationally complete.
This result gives an additional aspect to the common observation that quantum tomography is different in even and odd dimensions \cite{ShVo11}.
\section{Preliminaries}\label{sec:preli}
In this section we fix some notations and introduce the basic concepts.
\subsection*{States and Observables}
Let $\hi$ be a finite dimensional Hilbert space, with $\dim\hi = d \geq 2$.
We denote by $\lh$ the vector space of all linear operators on $\hi$.
A positive operator $\varrho\in\lh$ having trace one is a \emph{state}, and we denote by $\sh$ the set of all states.
Observables are generally described by \emph{positive operator valued measures} (POVMs) \cite{OQP97,FQMEA02}.
In this work we only consider observables with finite number of outcomes.
Therefore, an observable can be defined as a function $\Ao:x\mapsto\Ao(x)$, where each $\Ao(x)$ is a positive operator and $\sum_x \Ao(x) = \id$.
Here the sum runs over all $x\in\Omega_\Ao$, where the set $\Omega_\Ao$ is the collection of all possible measurement outcomes.
If a system is prepared in a state $\varrho$, then a measurement of an observable $\Ao$ will lead to an outcome $x$ with the probability $\tr{\varrho\Ao(x)}$.
\subsection*{Informational completeness}
An observable $\Ao$ is \emph{informationally complete} if its measurement outcome probability distribution is sufficient to identify a unique state \cite{Prugovecki77}.
In other words, two different states must give rise to different probability distributions: for all pairs of states $\varrho_1,\varrho_2$,
\begin{equation*}
\tr{\varrho_1 \Ao(j)} = \tr{\varrho_2 \Ao(j)} \quad \forall j\in\Omega_\Ao \quad \Rightarrow \quad\varrho_1=\varrho_2 \, .
\end{equation*}
The informational completeness of an observable $\Ao$ is equivalent to the property that the linear span of the set $\{\Ao(j):j\in\Omega_\Ao\}$ is $\lh$ \cite{Busch91,SiSt92}.
\subsection*{Joint measurability}
Given two observables $\Ao$ and $\Bo$, we say that they are \emph{jointly measurable} if there exists a third observable $\Co$ with $\Omega_\Co=\Omega_\Ao \times\Omega_\Bo$ and satisfying
\begin{equation}\label{eq:joint-definition}
\sum_{x\in\Omega_\Ao}\Co(x,y)=\Bo(y) \quad \forall y \, , \qquad \sum_{y\in\Omega_\Bo} \Co(x,y)=\Ao(x) \quad \forall x \, .
\end{equation}
In other words, $\Ao$ and $\Bo$ correspond to the `marginals' of $\Co$.
Any observable satisfying \eqref{eq:joint-definition} is called a \emph{joint observable} of $\Ao$ and $\Bo$ \cite{LaPu97}.
We recall that joint measurability is equivalent to the following \cite{AlCaHeTo09}: there exists an observable $\Go$ and stochastic matrices $[M_{xz}]$, $[M'_{yz}]$ such that
\begin{equation}
\sum_z M_{xz} \Go(z)=\Ao(x) \quad \forall x\, , \quad \sum_z M'_{yz} \Go(z)=\Bo(y) \quad \forall y \, .
\end{equation}
Hence, two observables are jointly measurable iff they can be `post-processed' from a single observable.
We will use several times the following simple fact: if $\Co$ and $\Co'$ are joint observables of $\Ao$ and $\Bo$, then also all their convex combinations $t\Co+(1-t)\Co'$, $0<t<1$, are joint observables of $\Ao$ and $\Bo$.
It follows that two jointly measurable observables have either a unique joint observable or infinitely many of them.
Another useful fact is related to unitary transformations.
Let $U$ be a unitary operator on $\hi$.
Two observables $\Ao$ and $\Bo$ are jointly measurable if and only if the observables $U\Ao U^\ast$ and $U \Bo U^\ast$ are jointly measurable.
Indeed, it is easy to see that $\Co$ is a joint observable of $\Ao$ and $\Bo$ if and only if $U\Co U^\ast$ is a joint observable of $U\Ao U^\ast$ and $U\Bo U^\ast$.
\subsection*{Instruments}
An observable describes the statistics of the outcomes of a measurement but leaves open how the measurement disturbs the input state.
In order to discuss this we need the concept of an \emph{instrument} \cite{QTOS76}.
An instrument with finitely many outcomes is a mapping $\mathcal{I}:x\mapsto\mathcal{I}_x$ such that each $\mathcal{I}_x$ is
a completely positive linear map on $\lh$ and $\sum_x \tr{\mathcal{I}_x(\varrho)}=1$ for all states $\varrho$.
The \emph{adjoint map} $\mathcal{I}_x^\ast$ of $\mathcal{I}_x$ is defined via the usual trace duality
\begin{equation*}
\tr{S \mathcal{I}_x(T)}=\tr{\mathcal{I}_x^\ast(S)T} \qquad \forall S,T\in\lh \, .
\end{equation*}
In other words, $\mathcal{I}_x^\ast$ and $\mathcal{I}_x$ correspond to the Heisenberg and Schr\"odinger pictures, respectively.
Suppose that $\Ao$ is an observable.
Then we say that an instrument $\mathcal{I}$ is \emph{$\Ao$-compatible} if $\mathcal{I}_x^\ast(\id)=\Ao(x)$ for every $x$.
Every $\Ao$-compatible instrument describes some particular kind of measurement of $\Ao$ \cite{Ozawa84}.
An example of an $\Ao$-compatible instrument is the \emph{L\"uders instrument} $\mathcal{I}^L$, defined by
\begin{equation*}
\mathcal{I}^L_x(\varrho)=\sqrt{\Ao(x)} \varrho \sqrt{\Ao(x)} \, .
\end{equation*}
Any other $\Ao$-compatible instrument $\mathcal{I}$ is of the form
\begin{equation*}
\mathcal{I}_x(\varrho) = \Ec_x \left( \mathcal{I}^L_x(\varrho) \right)
\end{equation*}
for some collection $\{\Ec_x\}$ of completely positive trace preserving maps on $\lh$ \cite{QI06}.
\subsection*{Sequential measurements}
By a \emph{sequential measurement} we mean a setting where two measurements are combined
into a third measurement by performing them one after the other \cite{DaLe70}.
Generally, the order in which the measurements are performed is crucial \cite{BuCaLa90}.
Suppose we have two $N$-outcome observables $\Ao,\Bo$ and we measure them subsequently; first $\Ao$ and then $\Bo$.
As a result, we have in total $N^2$ possible measurement outcomes.
Generally, we do not obtain a joint measurement of $\Ao$ and $\Bo$ since the first measurement distrubs the input state.
In fact, the overall measurement depends on the way we measure $\Ao$.
If the first measurement is described by an $\Ao$-compatible instrument $\mathcal{I}$, then
the overall observable $\Co$ is given by
\begin{equation*}
\tr{\varrho \Co(j,k)} = \tr{\Bo(k)\mathcal{I}_j(\varrho)}
\end{equation*}
for all input states $\varrho$, or equivalently,
\begin{equation*}
\Co(j,k) = \mathcal{I}_j^\ast(\Bo(k)) \, .
\end{equation*}
Let us notice that first marginal of $\Co$ is always $\Ao$, while the second marginal is a perturbed version of $\Bo$ and depends on the instrument $\mathcal{I}$.
\section{Example: sequential measurements of $\sigma_x$ and $\sigma_y$}\label{sec:qubit}
We start with a preliminary example, which is mainly a collection of well known facts.
It hints the forthcoming developments and clarifies the aims of the later sections.
We refer to \cite{BuHe08} for more details and further references.
Fix $\hi=\C^2$, and let $\Ao$ and $\Bo$ be the two observables corresponding to the measurements of spin-$\half$ components in the directions $x$ and $y$, respectively.
Thus,
\begin{equation*}
\Ao(\pm 1) = \half (\id\pm\sigma_x) \, , \quad \Bo(\pm 1) = \half (\id\pm\sigma_y) \, ,
\end{equation*}
where $\sigma_x,\sigma_y$ are the Pauli spin matrices.
Since $\Ao$ and $\Bo$ consist of projections and they do not mutually commute, it is not possible to measure them jointly.
Moreover, if we measure them separately on two similarly prepared ensembles, we still cannot infer the unknown state.
An alternative way is to perform a sequential measurement.
The first measurement has to be a weak measurement, meaning that we do not measure $\Ao$ but its unsharp version.
We define an unsharp version $\Ao_\lambda$ of $\Ao$ by
\begin{equation*}
\Ao_\lambda(j) := \lambda \Ao(j) + (1-\lambda) \frac{1}{2} \id \, , \quad j=\pm 1 \, .
\end{equation*}
Here $\lambda\in[0,1]$ is a parameter quantifying the noise or imprecision.
We can write $\Ao_\lambda$ in the form
\begin{equation*}
\Ao_\lambda(\pm 1)= \half \bigl( \id \pm \lambda \sigma_x \bigr) \, .
\end{equation*}
In a similar way we define an unsharp version $\Bo_\gamma$ of $\Bo$ by
\begin{equation*}
\Bo_\gamma(\pm 1)= \half \bigl( \id \pm \gamma \sigma_y \bigr) \, .
\end{equation*}
We want to study the disturbance of the first measurement on the system, and for this reason we define an instrument related to $\Ao_\lambda$.
A class of $\Ao_\lambda$-compatible instruments can be defined by
\begin{equation*}
\mathcal{J}_{\pm 1}(\varrho) = L_{\pm 1}\sqrt{\Ao_\lambda(\pm 1)}\varrho\sqrt{\Ao_\lambda(\pm 1)}L_{\pm 1}^\ast \, ,
\end{equation*}
where $L_1,L_{-1}$ are arbitrary unitary operators.
If the subsequent measurement is a $\Bo$-measurement, then the overall statistics of the sequential measurement is given by the observable
\begin{equation*}
\Co(j,k) = \mathcal{J}_j^\ast(\Bo(k)) = \sqrt{\Ao_\lambda(j)}L_j^\ast \Bo(k) L_j \sqrt{\Ao_\lambda(j)} \, , \quad j,k=\pm 1 \, .
\end{equation*}
The properties of $\Co$ obviously depend on $L_1$ and $L_{-1}$.
In the following we consider two different choices of $L_{\pm 1}$.
\subsection*{Optimal joint measurement}
If we choose $L_{\pm 1}=\id$, then we obtain
\begin{equation*}
\Co(j,k) =\frac{1}{4} \bigl( \id + j\ \lambda \sigma_x + k\ \sqrt{1-\lambda^2} \sigma_y \bigr) \, , \quad j,k=\pm 1 \, .
\end{equation*}
In particular, the marginals are
\begin{eqnarray*}
\Co(j,+1) + \Co(j,-1) &=& \Ao_\lambda(j) \, , \\
\Co(+1,k) + \Co(-1,k) &=& \frac{1}{2} \bigl( \id + k\ \sqrt{1-\lambda^2}\ \sigma_y \bigr) = \Bo_{\sqrt{1-\lambda^2}}(k) \, .\\
\end{eqnarray*}
The joint observable $\Co$ is an optimal approximate joint measurement of $\sigma_x$ and $\sigma_y$.
This means that the unsharp parameters $\lambda$ and $\gamma=\sqrt{1-\lambda^2}$ saturate the inequality
\begin{equation}\label{eq:ur-qubit-2}
\lambda^2 + \gamma^2 \leq 1 \, .
\end{equation}
Indeed, it is known that this inequality is a necessary and sufficient criterion for two observables $\Ao_\lambda$ and $\Bo_\gamma$ to be jointly measurable \cite{Busch86}.
Let us also notice that the joint observable of $\Ao_\lambda$ and $\Bo_\gamma$ is unique if $\lambda^2 + \gamma^2 = 1$ \cite{Busch86}.
\subsection*{Informationally complete joint measurement}
Another interesting option is to choose
\begin{equation*}
L_{\pm 1} = \cos\frac{\theta}{2}\id \mp i \sin\frac{\theta}{2}\sigma_x
\end{equation*}
for some fixed angle $0<\theta < \pi/2$.
In this case we obtain
\begin{align*}
\Co(j,k) &= \frac{1}{4} \bigl( \id +j \ \lambda\sigma_x + k \ \cos\theta\sqrt{1-\lambda^2} \sigma_y + jk \ \sin\theta\sqrt{1-\lambda^2} \sigma_z\bigr)
\end{align*}
and the marginals are
\begin{eqnarray*}
\Co(j,+1) + \Co(j,-1) &=& \Ao_\lambda(j) \, , \\
\Co(+1,k) + \Co(-1,k) &=& \frac{1}{2} \bigl( \id + k \ \cos\theta\sqrt{1-\lambda^2} \sigma_y \bigr) = \Bo_{\cos\theta\sqrt{1-\lambda^2}}(k) \, .
\end{eqnarray*}
It is easy to see that the linear span of the four operators $\Co(j,k)$, $j,k=\pm1$, is the set of all $2\times 2$ - complex matrices.
It follows that the joint observable $\Co$ is informationally complete.
The unsharpness parameters $\lambda$ and $\gamma=\cos\theta\sqrt{1-\lambda^2}$ do not saturate the inequality \eqref{eq:ur-qubit-2}.
Altering the parameter $\theta$ we can make the sum $\lambda^2+\gamma^2$ as close to $1$ as we want, hence we conclude that
$\Ao_\lambda$ and $\Bo_\gamma$ admit an informationally complete joint observable if and only if
\begin{equation}
\lambda^2 + \gamma^2 < 1 \, .
\end{equation}
Finally, we remark that with the choices $\lambda=1/\sqrt{3}$ and $\theta=\pi/4$ the joint observable $\Co$ is a symmetric informationally complete (SIC) observable.
\section{Conjugate observables}\label{sec:conjugate}
\subsection{Mutually unbiased bases and complementary observables}\label{sec:mub}
We start by recalling the usual definition of complementary observables in a finite $d$-dimensional Hilbert space and some related basic facts \cite{Schwinger60}, \cite{Kraus87}.
We denote $\Z_d\equiv \{0,\ldots,d-1\}$.
Let $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$ be \emph{mutually unbiased bases} (MUBs), i.e., they are orthonormal bases in $\hi$ and
\begin{equation}\label{eq:mub}
\mo{\ip{\varphi_j}{\psi_k}}^2 = 1/d \qquad \forall j,k \in\Z_d \, .
\end{equation}
We define two $d$-outcome observables $\Ao$ and $\Bo$ corresponding to $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$, respectively.
Hence,
\begin{equation}\label{eq:AB}
\Ao(j)=\kb{\varphi_j}{\varphi_j} \, , \qquad \Bo(k)=\kb{\psi_k}{\psi_k} \, .
\end{equation}
Obviously, two orthonormal bases $\{\varphi_j\}_{j\in\Z_d}$ and $\{\varphi'_j\}_{j\in\Z_d}$ define the same observable $\Ao$ iff $\varphi'_j=\alpha_j \varphi_j$ for some complex numbers $\alpha_j$ of modulus one.
The mutual unbiasedness condition \eqref{eq:mub} can be rephrased by saying that $\Ao$ and $\Bo$ are \emph{complementary} observables, meaning that in any state $\varrho$ where the outcome of $\Ao$ is predictable, the $\Bo$-distribution is uniform (and vice versa).
This entails that the following implications are valid for any state $\varrho$ and all outcomes $j,k\in\Z_d$,
\begin{align*}
\tr{\varrho\Ao(j)} &= 1 \quad\Rightarrow\quad \tr{\varrho\Bo(k)} = 1/d \\
\tr{\varrho\Bo(k)} &= 1 \quad\Rightarrow\quad \tr{\varrho\Ao(j)} = 1/d \, .
\end{align*}
Since $\tr{\varrho\Ao(j)} = 1$ iff $\varrho=\kb{\varphi_j}{\varphi_j}$, it is easy to see that the complementarity of $\Ao$ and $\Bo$ is indeed equivalent to the mutual unbiasedness of the bases $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$.
There is a canonical way to produce two mutually unbiased bases.
In the following, suppose an orthonormal basis $\{\varphi_k\}_{k\in\Z_d}$ of $\hh$ is fixed.
Denoting $\omega\equiv e^{2\pi i /d}$, we define the following unitary representations $U$ and $V$ of the cyclic group $\Z_d$ in $\hh$:
\begin{eqnarray*}
U_x \varphi_k & := & \varphi_{k+x} \\
V_y \varphi_k & := & \omega^{yk} \varphi_k
\end{eqnarray*}
for all $x,y,k \in\Z_d$.
In the above formulas and in the rest of the paper, addition and multiplication of elements in $\Z_d$ are understood modulo $d$.
(For instance, we will often use $-j=d-j$).
It is easy to verify that
\begin{equation}\label{eq:VU}
V_y U_x = \omega^{xy} \ U_x V_y \qquad \forall x,y\in\Z_d \, .
\end{equation}
The Fourier transform (with respect to the basis $\{\varphi_k\}_{k\in\Z_d}$) is the unitary operator $\mathcal{F} : \hh \to \hh$ defined by
\begin{equation}\label{eq:defF}
\mathcal{F} \varphi_k := \frac{1}{\sqrt{d}} \sum_{h\in\Z_d} \omega^{-hk} \varphi_h \, .
\end{equation}
The adjoint operator $\mathcal{F}^\ast$ of $\mathcal{F}$ is given by
\begin{equation}\label{eq:defF-adjoint}
\mathcal{F}^\ast \varphi_k = \frac{1}{\sqrt{d}} \sum_{h\in\Z_d} \omega^{hk} \varphi_h = \mathcal{F}\varphi_{-k} \, ,
\end{equation}
and we have $\mathcal{F}^2 \varphi_k = \mathcal{F}^{\ast \, 2} \varphi_k = \varphi_{-k}$.
We denote
\begin{equation*}
\psi_k \equiv \mathcal{F}^\ast \varphi_k = \mathcal{F} \varphi_{-k} \, ,
\end{equation*}
and it is immediate to check that $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$ are MUBs, with $\ip{\varphi_j}{\psi_k} = (1/\sqrt{d}) \, \omega^{jk}$.
The Fourier transform has the intertwining properties
$$
\mathcal{F} U_x = V_x^\ast \mathcal{F} \, , \qquad \mathcal{F} V_y = U_y \mathcal{F} \, ,
$$
from which it follows that
\begin{eqnarray*}
U_x \psi_k & = & \omega^{-xk} \psi_k \\
V_y \psi_k & = & \psi_{k+y} \, .
\end{eqnarray*}
The observables $\Ao$ and $\Bo$ related to $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$, respectively, satisfy the following conditions for all $j,k,x,y\in\Z_d$:
\begin{equation}\label{eq:cov-A}
U_x \Ao(j) U_x^\ast = \Ao(j+x) \, , \quad V_y \Ao(j) V_y^\ast = \Ao(j)
\end{equation}
and
\begin{equation}\label{eq:cov-B}
U_x \Bo(k) U_x^\ast = \Bo(k) \, , \quad V_y \Bo(k) V_y^\ast = \Bo(k+y) \, .
\end{equation}
In other words, $\Ao$ is $U$-covariant and $V$-invariant, while $\Bo$ is $U$-invariant and $V$-covariant.
We also note that $\Ao$ and $\Bo$ are conjugated by $\mathcal{F}$, i.e.,
\begin{equation}
\Bo(k) = \mathcal{F}^\ast \Ao(k) \mathcal{F}
\end{equation}
for all $k\in\Z_d$.
It is customary to say that $\Ao$ and $\Bo$ are {\em canonically conjugated} observables.
The conditions \eqref{eq:cov-A} -- \eqref{eq:cov-B} are analogous to the symmetry properties of the usual position and momentum observables on the real line $\R$ (see e.g. \cite{CaHeTo04}).
In some situations the covariance properties may have some physical meaning or motivation.
However, for our purposes they are just useful features that can be utilized later in our calculations.
\subsection{Unsharp observables}\label{sec:unsharp}
Measurements of two complementary observables are incompatible and therefore have to be performed separately.
This means that their measurements require different settings.
However, it is possible to perform a simultaneous measurement of two complementary observables if we allow some additional imprecision or noise.
In other words, we can measure jointly \emph{unsharp versions} of $\Ao$ and $\Bo$.
We define an unsharp version $\Ao_\lambda$ of $\Ao$ by
\begin{equation*}
\Ao_\lambda(j) := \lambda \Ao(j) + (1-\lambda) \frac{1}{d} \id \, .
\end{equation*}
Here $\lambda\in[0,1]$ is a parameter quantifying the noise.
This type of noise is equivalent to the situation where an input state $\varrho$ is first depolarized into a state $\lambda \varrho + (1-\lambda)/d \id$ and then a measurement of $\Ao$ is performed.
More generally, if $\Lambda$ is a probability distribution on $\Z_d$, then we define an unsharp version $\Ao_\Lambda$ of $\Ao$ by
\begin{equation}\label{eq:defALam}
\Ao_\Lambda(j) := \sum_{i\in\Z_d} \Lambda(j-i) \Ao(i) \, .
\end{equation}
The special case $\Ao_\Lambda=\Ao_\lambda$ corresponds to the probability distribution $\Lambda$ defined as
\begin{equation*}
\Lambda(0) = \lambda + (1-\lambda)/d \, , \quad \Lambda(j) = (1-\lambda)/d \quad \textrm{if}\ j\neq 0 \, .
\end{equation*}
We can also write this probability distribution in the form
\begin{equation*}
\Lambda(j) = \lambda \delta (j) + (1-\lambda) \mu (j)
\end{equation*}
where $\delta$ is the point distribution at $0$ and $\mu$ is the uniform distribution on $\Z_d$, i.e.,
\begin{equation}\label{eq:point-and-uniform}
\delta (j) = \left\{\begin{array}{cc} 1 & \quad \mbox{if } j=0 \\ 0 & \quad \mbox{if } j\neq 0 \end{array} \right. \, , \qquad \mu (j) = \frac{1}{d} \quad \forall j \, .
\end{equation}
In a similar way a probability distribution $\Gamma$ on $\Z_d$ defines an unsharp version $\Bo_\Gamma$ of $\Bo$ by
\begin{equation}\label{eq:defBGamma}
\Bo_\Gamma(k) := \sum_{i\in\Z_d} \Gamma(k-i) \Bo(i) \, .
\end{equation}
A special class is, again, characterized by noise parameters $\gamma \in [0,1]$ and we denote
\begin{equation*}
\Bo_\gamma(k) := \gamma \Bo(k) + (1-\gamma) \frac{1}{d} \id \, .
\end{equation*}
Naturally, there are also other type of approximations of $\Ao$ and $\Bo$ than the previously defined $\Ao_\Lambda$ and $\Bo_\Gamma$.
The usefulness of $\Ao_\Lambda$ and $\Bo_\Gamma$ is that they satisfy the same covariance and invariance relations than $\Ao$ and $\Bo$, respectively.
Namely, the observables $\Ao_\Lambda$ and $\Bo_\Gamma$ satisfy the following conditions:
\begin{equation}\label{eq:cov-A-unsharp}
U_x \Ao_\Lambda(j) U_x^\ast = \Ao_\Lambda(j+x) \, , \quad V_y \Ao_\Lambda(j) V_y^\ast = \Ao_\Lambda(j)
\end{equation}
and
\begin{equation}\label{eq:cov-B-unsharp}
U_x \Bo_\Gamma(k) U_x^\ast = \Bo_\Gamma(k) \, , \quad V_y \Bo_\Gamma(k) V_y^\ast = \Bo_\Gamma(k+y) \, .
\end{equation}
Thus, $\Ao_\Lambda$ and $\Bo_\Gamma$ are \emph{conjugated observables} although they need not be complementary anymore \cite{CaHeTo11}.
As we will see, two observables $\Ao_\Lambda$ and $\Bo_\Gamma$ can have a joint observable even if they do not commute.
\begin{remark}
Suppose that $\widetilde{\Ao}$ is a $d$-outcome observable satisfying
\begin{equation}\label{eq:cov-A-general}
U_x \widetilde{\Ao}(j) U_x^\ast = \widetilde{\Ao}(j+x) \, , \quad V_y \widetilde{\Ao}(j) V_y^\ast = \widetilde{\Ao}(j)
\end{equation}
for all $j,x,y\in\Z_d$.
Then $\widetilde{\Ao}=\Ao_\Lambda$ for some probability distribution $\Lambda$.
Namely, it follows from the second condition in \eqref{eq:cov-A-general} that $\widetilde{\Ao}$ commutes with $\Ao$ (since $\Ao(j) = (1/d) \sum_y \omega^{-jy} V_y$) and hence $\widetilde{\Ao}(j)=\sum_k p_{j,k} \Ao(k)$ for some real numbers $0\leq p_{j,k} \leq 1$.
The first condition in \eqref{eq:cov-A-general} then implies that $p_{j,k}=p_{0,k-j}$.
\end{remark}
\section{Joint measurements}\label{sec:joint}
\subsection{Covariant observables}
We recall that two observables $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable if they have a joint observable, i.e., an observable $\Co$ on $\Z_d\times\Z_d$ such that
\begin{equation}\label{eq:joint-povm}
\sum_{k\in\Z_d} \Co(j,k) = \Ao_\Lambda(j) \, , \qquad \sum_{j\in\Z_d} \Co(j,k) = \Bo_\Gamma(k)
\end{equation}
for all $j,k\in\Z_d$.
A special class of joint observables turns out to be crucial for our developments.
We say that an observable $\Co$ on $\Z_d\times\Z_d$ is a \emph{covariant phase space observable} if
\begin{equation}
U_xV_y\Co(j,k)V_y^\ast U_x^\ast = \Co(j+x,k+y)
\end{equation}
for all $j,k,x,y\in\Z_d$.
The covariant phase space observables have a simple form \cite{QTOS76}.
Namely, if $\Co$ is a covariant phase space observable, then there is unique operator $T\in\sh$ such that $\Co=\Co_T$, where we have denoted
\begin{equation*}
\Co_T(j,k):=\frac{1}{d}\ U_jV_k T V_k^\ast U_j^\ast \, , \qquad j,k\in\Z_d \, .
\end{equation*}
Also, each $T\in\sh$ defines a covariant phase space observable by this formula.
The correspondence $T\leftrightarrow \Co_T$ is therefore one-to-one and the elements in $\sh$ parametrize the covariant phase space observables.
The marginals of a covariant phase space observable $\Co_T$ are conjugated observables on $\Z_d$.
Indeed, a direct calculation shows that $\Co_T$ has marginals $\Ao_\Lambda$ and $\Bo_\Gamma$, with
\begin{equation}\label{eq:structure1}
\Lambda (j) = \tr{\Ao (-j)T} \, , \qquad
\Gamma (k) = \tr{\Bo (-k)T} \, .
\end{equation}
(This calculation can be found in \cite{CaHeTo11}).
The essential role of covariant phase space observables in our discussion becomes clear in the following observation.
\begin{proposition}\label{prop:if-jm-then-cov}
If $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable, then they have a joint observable which is a covariant phase space observable.
\end{proposition}
\begin{proof}
Suppose that $\Co$ is a joint observable of $\Ao_\Lambda$ and $\Bo_\Gamma$.
For each $x,y\in\Z_d$, we define an observable $\Co_{x,y}$ by
\begin{equation}\label{eq:cxy}
\Co_{x,y}(j,k) := U_x^\ast V_y^\ast \Co(j+x,k+y) V_y U_x \, , \qquad j,k\in\Z_d \, .
\end{equation}
Using the covariance and invariance properties \eqref{eq:cov-A-unsharp} -- \eqref{eq:cov-B-unsharp} it is straightforward to verify that $\Co_{x,y}$ is a joint observable of $\Ao_\Lambda$ and $\Bo_\Gamma$.
We then define $\widetilde{\Co}$ to be the uniform mixture of all $\Co_{x,y}$, i.e.,
\begin{equation}\label{eq:ctilde}
\widetilde{\Co}(j,k) := \frac{1}{d^2}\sum_{x,y\in\Z_d} \Co_{x,y}(j,k) \, , \qquad j,k\in\Z_d \, .
\end{equation}
Since every $\Co_{x,y}$ is a joint observable of $\Ao_\Lambda$ and $\Bo_\Gamma$, also $\widetilde{\Co}$ is their joint observable.
A direct calculation, using \eqref{eq:VU}, shows that $\widetilde{\Co}$ is a covariant phase space observable.
\end{proof}
We conclude from Proposition \ref{prop:if-jm-then-cov} that two observables $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable iff their related probability distributions $\Lambda$ and $\Gamma$ are of the form \eqref{eq:structure1} for some $T\in\sh$. Equations \eqref{eq:structure1} can also be rewritten in a slightly different form. Namely, observe that, if $T\in\sh$, then there exists a unit vector $\phi\in\hh\otimes\hh$ such that $T={\rm tr}_2 [\kb{\phi}{\phi}]$, where ${\rm tr}_2$ is the partial trace with respect to the second factor.
(A vector state giving $T$ via the partial trace is often called a \emph{purification} of $T$).
Conversely, if $\phi\in\hh\otimes\hh$ is a unit vector, then $T={\rm tr}_2 [\kb{\phi}{\phi}]$ is a state.
Inserting this form into \eqref{eq:structure1} we obtain
\begin{equation}\label{eq:structure2}
\Lambda (j) = \ip{\phi}{(\Ao (-j)\otimes \id)\phi} \, , \quad \Gamma (k) = \ip{\phi}{(\Bo(-k)\otimes \id)\phi} \, .
\end{equation}
Note that if a vector $\phi\in\hh\otimes\hh$ satisfies the above two equations for some probability densities $\Lambda$ and $\Gamma$, then the normalization $\no{\phi} = 1$ is automatic.
We thus have the following characterization of jointly measurable observables.
\begin{proposition}\label{prop:joint->phi}
Let $\Lambda$, $\Gamma$ be probability densities on $\Z_d$. The following facts are equivalent:
\begin{itemize}
\item[{\rm (i)}] The observables $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable.
\item[{\rm (ii)}] There exists a state $T \in \sh$ such that the probability densities $\Lambda$ and $\Gamma$ satisfy \eqref{eq:structure1}.
\item[{\rm (iii)}] There exists a vector $\phi \in \hh\otimes\hh$ such that the probability densities $\Lambda$ and $\Gamma$ satisfy \eqref{eq:structure2}.
\end{itemize}
\end{proposition}
Let us note that any $\Ao_\Lambda$ is jointly measurable with some $\Bo_\Gamma$.
Namely, for each $\Lambda$ we can define a state $T_\Lambda$ as
\begin{equation}
T_\Lambda := \sum_{j\in\Z_d} \Lambda(-j) \kb{\varphi_j}{\varphi_j} \, .
\end{equation}
Then $\Lambda (j) = \tr{\Ao (-j)T_\Lambda}$ and $\Ao_\Lambda$ is thus jointly measurable with $\Bo_\Gamma$, where $\Gamma$ is defined as $\Gamma (k) = \tr{\Bo (-k)T_\Lambda}$.
Proposition \ref{prop:joint->phi} can be seen as a trade-off relation between the probability distributions $\Lambda$ and $\Gamma$ that describe the deviations of $\Ao_\Lambda$ and $\Bo_\Gamma$ from $\Ao$ and $\Bo$, respectively.
For instance, if $\Lambda=\delta$, then necessarily $\Gamma=\mu$.
Hence, we recover the fact that $\Ao$ is jointly measurable only with the trivial observable and no other $\Bo_\Gamma$.
We end this subsection with some additional observations.
\begin{remark}
\emph{If two observables $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable, they can have several different covariant phase space observables as their joint observables.}
For instance, let $\{\zeta_i\}_{i\in\Z_d}$ be an orthonormal basis which is mutually unbiased with respect to both orthonormal bases $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$.
Then, for each $i\in\Z_d$, we have
\begin{equation*}
\ip{\zeta_i}{\Ao (-j)\zeta_i}=\ip{\zeta_i}{\Bo (-k)\zeta_i}=1/d \, .
\end{equation*}
Therefore, the marginals of the covariant phase space observables $\Co_{\kb{\zeta_i}{\zeta_i}}$ are the same although $\Co_{\kb{\zeta_i}{\zeta_i}}\neq\Co_{\kb{\zeta_{i'}}{\zeta_{i'}}}$ whenever $i\neq i'$.
\end{remark}
\begin{remark}
\emph{If two observables $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable, they can have a joint observable which is not a covariant phase space observable.}
For instance, let $p:\Z_d\times\Z_d\to[0,1]$ be a bivariate probability distribution with uniform margins.
Then the observable $\Co(j,k):=p(j,k)\id$ is a joint observable of $\Ao_0$ and $\Bo_0$.
It is clear that $\Co$ is a covariant phase space observable only if $p$ is a uniform distribution.
However, a bivariate probability distribution with uniform margins need not be uniform.
For instance, if we set
\begin{equation*}
p(i,j) = \frac{1}{d^2} \left( 1 - \sin \left( 2\pi \frac{ij}{d} \right) \right) \, , \quad i,j\in\Z_d \, ,
\end{equation*}
then $\sum_i p(i,j)=\sum_j p(i,j)=1/d$, but $p$ is not uniform.
The existence of non-covariant joint observables is not limited to the trivial observables $\Ao_0$ and $\Bo_0$.
Namely, suppose that $\Co$ is a joint observable of $\Ao_0$ and $\Bo_0$ and $\Co'$ is a joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.
Then the convex combination $t\Co' + (1-t)\Co$, $0<t<1$, is a joint observable of $\Ao_{t\lambda}$ and $\Bo_{t\gamma}$.
It is easy to see that if $\Co'$ is a covariant phase space observable but $\Co$ is not, then their convex combination $t\Co' + (1-t)\Co$ cannot be a covariant phase space observable.
\end{remark}
\begin{remark}\label{prop:unique}
\emph{Suppose that $\Ao_\lambda$ and $\Bo_\gamma$ have a unique joint observable $\Co_T$ among the covariant phase space observables and that $T^2=T$.
Then $\Co_T$ is a unique joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.}
To prove this claim, let $\Co$ be a joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.
We need to show that $\Co=\Co_T$.
We define $\Co_{x,y}$ and $\widetilde{\Co}$ as in \eqref{eq:cxy} -- \eqref{eq:ctilde}.
Since $\widetilde{\Co}$ is by its construction a covariant phase space observable, we must have $\widetilde{\Co}=\Co_T$ by the assumption on uniqueness.
In particular, each operator $\widetilde{\Co}(j,k)=(1/d)\, U_jV_k T V_k^\ast U_j^\ast$ is rank-1. Since $0\leq \Co_{x,y} (0,0) \leq d^2 \, \widetilde{\Co}(0,0) = d\, T$ by \eqref{eq:ctilde}, it follows that there exists a real constant $0\leq c(x,y) \leq d$ such that $\Co_{x,y} (0,0) = c(x,y) \, T$, hence
$$
\Co(x,y) = c(x,y) \, U_x V_y T V_y^\ast U_x^\ast
$$
by \eqref{eq:cxy}.
Suppose $\lambda \neq 1$.
Since $\Co$ has $\Ao_\lambda$ as its first marginal, then
$$
\sum_{y\in\Z_d} c(x,y) \, U_x V_y T V_y^\ast U_x^\ast = \lambda \Ao (x) + (1-\lambda) \frac{1}{d} \id \quad \forall x\in\Z_d \, .
$$
The right hand side of this equation is a rank-$d$ operator, while on the left hand side we have the sum of $d$ operators with rank-$1$.
It then follows that the set $\{ U_x V_y T V_y^\ast U_x^\ast \}_{y\in\Z_d}$ is linearly independent in $\lh$. Since $\Co$ and $\widetilde{\Co}$ have the same marginals,
$$
\sum_{y\in\Z_d} c(x,y) \, U_x V_y T V_y^\ast U_x^\ast = \frac{1}{d} \sum_{y\in\Z_d} U_x V_y T V_y^\ast U_x^\ast
$$
for all $x$, hence $c(x,y) = 1/d$ for all $x,y$ by linear independence.
The case $\lambda = 1$ is treated in a similar way, by taking the marginal $\Bo_\gamma$ in the place of $\Ao_\lambda$ (and now necessarily $\gamma\neq 1$).
Therefore, $\Co_T$ is the unique joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.
\end{remark}
\subsection{Unsharpness inequality}
In this subsection we apply Proposition \ref{prop:joint->phi} to the cases where $\Ao_\Lambda=\Ao_\lambda$ and $\Bo_\Gamma=\Bo_\gamma$.
These special type of marginal observables are interesting as we can quantify their unsharpnesses by single numbers $\lambda$ and $\gamma$.
In particular, we can ask how small $\lambda$ and $\gamma$ must be in order for $\Ao_\lambda$ and $\Bo_\gamma$ to become jointly measurable.
We first notice that this question is, indeed, meaningful.
\begin{proposition}\label{prop:ord. makes sense}
Let $\lambda,\gamma\in (0,1]$. The following conditions are equivalent:
\begin{itemize}
\item[(i)] $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable.
\item[(ii)] $\Ao_{\lambda'}$ and $\Bo_{\gamma'}$ are jointly measurable for all $0\leq\lambda' \leq \lambda$ and $0\leq\gamma'\leq\gamma$.
\item[(iii)] $\Ao_{\lambda'}$ and $\Bo_{\gamma'}$ are jointly measurable for all $0\leq\lambda' <\lambda$ and $0\leq\gamma' <\gamma$.
\end{itemize}
\end{proposition}
\begin{proof}
Suppose that (i) holds and $0<\gamma' <\gamma$.
We denote $t:=\gamma'/\gamma$ and hence $0<t < 1$.
We have
\begin{equation*}
\Bo_{\gamma'}(k) = t \Bo_{\gamma}(k) + (1-t) \frac{1}{d} \id \, ,
\end{equation*}
meaning that $\Bo_{\gamma'}$ is a convex combination of $\Bo_{\gamma}$ and the trivial observable $\Bo_0$.
By the assumption $\Ao_\lambda$ is jointly measurable with $\Bo_{\gamma}$ and $\Ao_\lambda$ is also jointly measurable with the trivial observable $\Bo_0$ (since they commute).
If $\Co_1$ is a joint observable of $\Ao_\lambda$ and $\Bo_{\gamma}$ and $\Co_2$ is a joint observable of $\Ao_\lambda$ and $\Bo_0$, then the convex combination $t\Co_1 + (1-t)\Co_2$ is a joint observable of $\Ao_\lambda$ and $\Bo_{\gamma'}$.
Therefore, $\Ao_\lambda$ and $\Bo_{\gamma'}$ are jointly measurable.
We can interchange the roles of $\Ao_\lambda$ and $\Bo_{\gamma}$ and run the same argument, hence we obtain (ii).
It is clear that (ii) implies (iii).
Hence, to complete the proof we need to show that (iii) implies (i).
Suppose that (iii) holds.
We choose sequences $(\lambda_n)$ and $(\gamma_n)$ such that $0<\lambda_n<\lambda$, $0<\gamma_n<\gamma$ and $\lim_n \lambda_n=\lambda$, $\lim_n\gamma_n=\gamma$.
For each $n$, we fix a state $T_n$ such that the corresponding covariant phase space observable $\Co_{T_n}$ is a joint observable of $\Ao_{\lambda_n}$ and $\Bo_{\gamma_n}$.
The set of states $\sh$ is compact in the operator norm topology, hence the sequence $(T_n)_n$ has a convergent subsequence.
We denote by $T$ the limit of this convergent subsequence.
Using \eqref{eq:structure1} we see that the covariant phase space observable $\Co_T$ is a joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.
Thus, (i) holds.
\end{proof}
We conclude from Proposition \ref{prop:ord. makes sense} that for every $\lambda\in[0,1]$, there is a number $\gamma_{\max}(\lambda) \geq 0$ such that $\Ao_{\lambda}$ and $\Bo_{\gamma}$ are jointly measurable iff $0\leq\gamma\leq\gamma_{\max}(\lambda)$.
Similarly, for every $\gamma\in[0,1]$, there is a number $\lambda_{\max}(\gamma)\geq 0$ such that $\Ao_{\lambda}$ and $\Bo_{\gamma}$ are jointly measurable iff $0\leq\lambda\leq\lambda_{\max}(\gamma)$.
We also know that $\gamma_{\max}(0)=\lambda_{\max}(0)=1$ (since a trivial observable is jointly measurable with any other observable) and that $\gamma_{\max}(1)=\lambda_{\max}(1)=0$ (see the discussion after Proposition \ref{prop:joint->phi}).
\begin{proposition}
The equality $\lambda_{\max} (x) = \gamma_{\max} (x)$ holds for all $x\in[0,1]$.
\end{proposition}
\begin{proof}
It is enough to show that $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable if and only if $\Ao_\gamma$ and $\Bo_\lambda$ are such.
Joint measurability of $\Ao_\lambda$ and $\Bo_\gamma$ means that there exists an observable $\Co$ on $\Z_d \times \Z_d$ having marginals $\Ao_\lambda$ and $\Bo_\gamma$, respectively.
We set $\widehat{\Co} (j,k) := \mathcal{F}^\ast \Co(k,-j) \mathcal{F}$ for every $j,k\in\Z_d$.
Then
\begin{align*}
\sum_{k\in\Z_d} \widehat{\Co} (j,k) & = \sum_{k\in\Z_d} \mathcal{F}^\ast \Co(k, -j) \mathcal{F} = \mathcal{F}^\ast \Bo_\gamma (-j) \mathcal{F} \\
& = \mathcal{F}^{\ast \, 2} \Ao_\gamma (-j) \mathcal{F}^2 = \Ao_\gamma (j) \, , \\
\sum_{j\in\Z_d} \widehat{\Co} (j, k) & = \sum_{j\in\Z_d} \mathcal{F}^\ast \Co(k , -j) \mathcal{F} = \mathcal{F}^\ast \Ao_\lambda (k) \mathcal{F} \\
& = \Bo_\lambda (k) \, .
\end{align*}
We conclude that $\widehat{\Co}$ is a joint observable of $\Ao_\gamma$ and $\Bo_\lambda$, hence the latter two are jointly measurable.
\end{proof}
We will now find out the function $\gamma_{\max}(\lambda)$, or, equivalently, $\lambda_{\max}(\gamma)$.
Suppose that $\lambda,\gamma\in [0,1]$ are such that $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable.
By Proposition \ref{prop:joint->phi} this is equivalent to the existence of a vector $\phi \in \hh\otimes\hh$ satisfying
\begin{eqnarray}
\ip{\phi}{(\Ao (j)\otimes \id)\phi} & = & \lambda \delta (j) + (1-\lambda) \mu (j) \, ,\label{eq. 1} \\
\ip{\phi}{(\Bo(k)\otimes \id)\phi} & = & \gamma \delta (k) + (1-\gamma) \mu (k)\, ,\label{eq. 2}
\end{eqnarray}
for all $j,k\in\Z_d$ (see \eqref{eq:point-and-uniform} for the definition of $\delta$ and $\mu$).
We now give a condition on the parameters $\lambda,\gamma$ which is necessary and sufficient for the existence of a vector $\phi \in \hh\otimes\hh$ satisfying the above two equations. Moreover, we show that, for the extreme values of $\lambda,\gamma$, the vector $\phi$ is essentially unique.
\begin{lemma}\label{lemma:bound}
Let $\lambda,\gamma\in [0,1]$.
Suppose there exists a vector $\phi\in\hh\otimes\hh$ satisfying \eqref{eq. 1} -- \eqref{eq. 2}.
Then
\begin{equation}\label{ineq. gamma}
\gamma \leq \frac{1}{d} \left[ (d-2)(1-\lambda) + 2 \sqrt{(1-d)\lambda^2 + (d-2)\lambda + 1} \right] \, .
\end{equation}
For any choice of a unit vector $\eta \in \hh$, the vector
\begin{equation}\label{phi est.}
\phi = (\alpha_\lambda \varphi_0 + \beta_\lambda \psi_0) \otimes \eta \, ,
\end{equation}
with
\begin{align*}
\alpha_\lambda & = \frac{1}{\sqrt{d}} \left[\sqrt{(d-1)\lambda + 1} - \sqrt{1-\lambda}\right] \, , \qquad
\beta_\lambda = \sqrt{1-\lambda} \, ,
\end{align*}
satisfies \eqref{eq. 1} -- \eqref{eq. 2} with equality in \eqref{ineq. gamma}.
Hence, the right hand side in \eqref{ineq. gamma} is equal to $\gamma_{\max}(\lambda)$.
If $\phi'$ is a vector satisfying \eqref{eq. 1} -- \eqref{eq. 2} with $\gamma=\gamma_{\max}(\lambda)$, then $\phi' = (\alpha_\lambda \varphi_0 + \beta_\lambda \psi_0) \otimes \eta'$ for some unit vector $\eta' \in \hh$.
\end{lemma}
\begin{proof}
Suppose $\phi\in\hh\otimes\hh$ satisfies \eqref{eq. 1} -- \eqref{eq. 2}.
We write $\phi$ in the form
$$
\phi = \sum_{i\in\Z_d} \varphi_i \otimes \xi_i \, ,
$$
where $\{ \xi_i\}_{i\in\Z_d}$ are vectors in $\hh$.
From \eqref{eq. 1} it follows that
\begin{equation*}
\no{\xi_i}^2 = \lambda \delta (i) + (1-\lambda) \mu (i) \, ,
\end{equation*}
hence there exist unit vectors $\{ \eta_i\}_{i\in\Z_d}$ such that
$$
\xi_0 = \sqrt{\frac{(d-1)\lambda +1}{d}} \, \eta_0 \, , \qquad \xi_i = \sqrt{\frac{1-\lambda}{d}} \, \eta_i \quad \forall i\neq 0 \, .
$$
On the other hand, we have
\begin{eqnarray*}
\ip{\phi}{(\Bo(k)\otimes \id)\phi} & = & \sum_{i,j\in\Z_d} \ip{\varphi_j\otimes\xi_j}{(\kb{\psi_k}{\psi_k} \otimes \id) \varphi_i\otimes\xi_i} \\
& = & \frac{1}{d} \sum_{i,j\in\Z_d} \omega^{jk} \omega^{-ik} \ip{\xi_j}{\xi_i} \, ,
\end{eqnarray*}
so, by \eqref{eq. 2}, we must have
$$
\frac{1}{d} \sum_{i,j\in\Z_d} \omega^{jk} \omega^{-ik} \ip{\xi_j}{\xi_i} = \gamma \delta (k) + (1-\gamma) \mu (k) \, .
$$
This equation, evaluated at $k=0$, gives
\begin{eqnarray*}
\left( 1 - \frac{1}{d} \right) \gamma + \frac{1}{d} & = & \frac{1}{d} \no{\sum_{i\in\Z_d} \xi_i}^2 \\
& = & \frac{1}{d^2} \no{\sqrt{(d-1)\lambda +1} \, \eta_0 + \sum_{i\in\Z_d, \, i\neq 0} \sqrt{1-\lambda} \, \eta_i}^2 \, .
\end{eqnarray*}
The maximum value of $\gamma$ is then achieved when the right hand side of this equation is maximal, i.e., when there exists a unit vector $\eta\in\hh$ such that $\eta_i = \eta$ $\forall i\in\Z_d$. The corresponding maximum value $\gamma_{\max}$ of $\gamma$ is given by
$$
\left( 1 - \frac{1}{d} \right) \gamma_{\max} + \frac{1}{d} = \frac{1}{d^2} \left( \sqrt{(d-1)\lambda +1} + (d-1) \sqrt{1-\lambda} \right)^2 \, ,
$$
i.e.,
\begin{equation*}
\gamma_{\max} = \frac{1}{d} \left[ (d-2)(1-\lambda) + 2 \sqrt{(1-d)\lambda^2 + (d-2)\lambda + 1} \right] \, .
\end{equation*}
In order to show that, if the sequence $\{\xi_i\}_{i\in\Z_d}$ is chosen as above, then the corresponding vector
\begin{eqnarray*}
\phi & = & \left( \sqrt{\frac{(d-1)\lambda +1}{d}} \, \varphi_0 + \sqrt{\frac{1-\lambda}{d}} \sum_{i\in\Z_d, \, i\neq 0} \varphi_i \right) \otimes \eta \\
& = & (\alpha_\lambda \varphi_0 + \beta_\lambda \psi_0) \otimes \eta
\end{eqnarray*}
satisfies also \eqref{eq. 2} with $\gamma=\gamma_{\max}$ (and thus the maximum is indeed achieved by $\phi$), we evaluate
\begin{eqnarray*}
\ip{\phi}{(\Bo(k)\otimes \id)\phi} & = & \ip{\phi}{(\kb{\psi_k}{\psi_k} \otimes \id)\phi} \\
& = & \left|\ip{\psi_k}{\alpha_\lambda \varphi_0 + \beta_\lambda \psi_0}\right|^2 \\
& = & \left( \frac{\alpha_\lambda}{\sqrt{d}} + \beta_\lambda \delta(k)\right)^2 \\
& = & \frac{\alpha^2_\lambda}{d} + \left(\beta^2_\lambda + 2 \frac{\alpha_\lambda \beta_\lambda}{\sqrt{d}} \right) \delta(k) \\
& = & (1-\gamma_{\max}) \mu (k) + \gamma_{\max} \delta (k) \, ,
\end{eqnarray*}
which is \eqref{eq. 2}.
\end{proof}
As a consequence of the above discussion, we obtain an inequality for the unsharpnesses of two jointly measurable observables $\Ao_\lambda$ and $\Bo_\gamma$.
\begin{figure}
\begin{center}
\includegraphics[width=4.0cm]{boundary.pdf}
\end{center}
\caption{The boundary curve $\lambda\mapsto\gamma_{\max}(\lambda)$ for $d=2,3,4,5$ (red color) and for $d=10,100,1000$ (orange color).}
\label{fig:boundary}
\end{figure}
\begin{proposition}\label{Prop:gammamax}
Two observables $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable if and only if
\begin{equation}\label{ineq. gamma 2}
\gamma \leq \gamma_{\max}(\lambda)=\frac{1}{d} \left[ (d-2)(1-\lambda) + 2 \sqrt{(1-d)\lambda^2 + (d-2)\lambda + 1} \right] \, ,
\end{equation}
(or, equivalently, its modified form under the exchange $\gamma\leftrightarrow\lambda$).
If $\gamma=\gamma_{\max}(\lambda)$, then $\Ao_\lambda$ and $\Bo_\gamma$ have a unique joint observable.
This unique joint observable is the covariant phase space observable $\Co_T$ defined by the state
$$
T = \kb{\chi_\lambda}{\chi_\lambda} \, , \qquad \chi_\lambda = \alpha_\lambda \varphi_0 + \beta_\lambda \psi_0 \, ,
$$
with
\begin{align}\label{eq:ab}
\alpha_\lambda &= \frac{1}{\sqrt{d}} \left[ \sqrt{(d-1) \lambda + 1} - \sqrt{1-\lambda} \right] \, , \qquad
\beta_\lambda = \sqrt{1-\lambda} \, .
\end{align}
\end{proposition}
\begin{proof}
If $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable, then the inequality follows from Proposition \ref{prop:joint->phi} and Lemma \ref{lemma:bound}.
Conversely, if $\gamma_{\max}$ is given by the right hand side of \eqref{ineq. gamma 2}, then the pair $\Ao_\lambda$ and $\Bo_{\gamma_{\max}}$ are jointly measurable again by an application of Proposition \ref{prop:joint->phi} and Lemma \ref{lemma:bound}.
Then, $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable by Proposition \ref{prop:ord. makes sense}.
Now suppose $\lambda$ and $\gamma$ achieve the bound \eqref{ineq. gamma 2}, and let $\Co_T$ be a covariant joint observable of $\Ao_\lambda$ and $\Bo_\gamma$.
Pick $\phi\in\hh\otimes\hh$ such that $T={\rm tr}_2 [\kb{\phi}{\phi}]$.
As $\phi$ satisfies \eqref{eq. 1} -- \eqref{eq. 2} with $\gamma = \gamma_{\max}$, by Lemma \ref{lemma:bound} it must be given by \eqref{phi est.} for some choice of a unit vector $\eta \in \hh$, and $T={\rm tr}_2 [\kb{\phi}{\phi}] = \kb{\chi_\lambda}{\chi_\lambda}$, with $\chi_\lambda$ as in the statement of the proposition.
Finally, we need to prove that $\Co_T$ is the unique joint observable (and not only unique among covariant phase space observables).
We notice that $T^2=T$, and the claim thus follows from Remark \ref{prop:unique}.
\end{proof}
The graph of the function $\lambda\mapsto\gamma_{\max}(\lambda)$ is a part of an ellipse.
In Fig.~\ref{fig:boundary} we have depicted it for $d=2,3,4,5,10,100,1000$.
\begin{example}
Suppose that two jointly measurable observables $\Ao_\lambda$ and $\Bo_\gamma$ are `equally unsharp' but as close to $\Ao$ and $\Bo$ as possible, i.e.,
\begin{equation*}
\gamma=\lambda = \lambda_{\max}(\gamma) \, .
\end{equation*}
In this case Proposition \ref{Prop:gammamax} gives
\begin{equation}
\gamma=\lambda = \frac{d+\sqrt{d}-2}{2(d-1)} = \half \left( 1+ \frac{1}{1+\sqrt{d}} \right) \, .
\end{equation}
The observables $\Ao_\lambda$ and $\Bo_\gamma$ then have a unique joint observable, which is the covariant phase space observable $\Co_T$ associated to the state $T=\kb{\chi}{\chi}$, with
$$
\chi =\sqrt{\frac{\sqrt{d}}{2(1+\sqrt{d})}} \, (\varphi_0 + \psi_0) \, .
$$
The vector state $\chi$ is hence an equal superposition of the vector states $\varphi_0$ and $\psi_0$.
\end{example}
By a direct calculation one can verify that the inequality \eqref{ineq. gamma 2} can be rewritten in the following equivalent form which is symmetric in $\lambda$ and $\gamma$.
\begin{proposition}\label{prop:symmetric}
Two observables $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable iff
\begin{equation}\label{eq:region}
\textrm{either} \quad \gamma+\lambda\leq 1 \quad \textrm{or} \quad \gamma^2 + \lambda^2 + \frac{2(d-2)}{d} (1- \gamma)(1- \lambda) \leq 1 \, .
\end{equation}
\end{proposition}
The second inequality in \eqref{eq:region} describes a full ellipse.
Therefore, the first condition $\gamma+\lambda\leq 1$ is needed to ignore the lower part of the ellipse, which is not a correct boundary for joint measurability.
This is depicted in Fig.~\ref{fig:ellipse}.
\begin{figure}
\begin{center}
\includegraphics[width=4.0cm]{ellipse.pdf}
\end{center}
\caption{The ellipse (red) and the line (dashed) from \eqref{eq:region} for $d=8$.
Only the upper side of the ellipse is relevant for the joint measurability.}
\label{fig:ellipse}
\end{figure}
Let us notice that for $d=2$ the latter inequality in \eqref{eq:region} becomes $\gamma^2 + \lambda^2\leq 1$, and then the first condition is redundant.
We thus recover the single condition stated in \eqref{eq:ur-qubit-2} and first proved in \cite{Busch86}.
Also, for $d=3$ and $d=4$ a direct calculation shows that the the first condition in \eqref{eq:region} is redundant and the latter quadratic inequality is necessary and sufficient for the joint measurability.
For $d\geq 5$ we need both conditions in \eqref{eq:region}.
By inspecting the function $\lambda\mapsto\gamma_{\max}(\lambda)$ we see that for every $\epsilon>0$, there is a pair of observables $\Ao_\lambda$ and $\Bo_\gamma$ such that they are not jointly measurable and $\lambda + \gamma < 1+ \epsilon$.
Thus, the criterion $\gamma+\lambda\leq 1$ is the best \emph{sufficient} condition for joint measurability which is linear and symmetric in $\lambda$ and $\gamma$.
The best linear and symmetric \emph{necessary} condition for joint measurablity is achived by taking the tangent of the boundary curve in the point where it crosses the line $\gamma=\lambda$.
In this way, we obtain the following conclusion.
\begin{proposition}\label{prop:linear}
If $\Ao_\lambda$ and $\Bo_\gamma$ are jointly measurable, then
\begin{equation}
\gamma+\lambda \leq 1+ \frac{\sqrt{d}-1}{d-1} \, .
\end{equation}
\end{proposition}
One can also see Proposition \ref{prop:linear} in the opposite order; if $\gamma+\lambda > 1+ \frac{\sqrt{d}-1}{d-1}$, then $\Ao_\lambda$ and $\Bo_\gamma$ are not jointly measurable.
In Fig.~\ref{fig:lines} we have depicted the linear necessary and sufficient conditions in the case $d=10$.
\begin{figure}
\begin{center}
\includegraphics[width=4.0cm]{lines.pdf}
\end{center}
\caption{In this picture $d=10$. The gray regions represent the necessary and sufficient linear conditions. In the white stripe one has to invoke the quadratic ellipse criterion, whereas otherwise the joint measurability can be deduced from the simple linear criteria.}
\label{fig:lines}
\end{figure}
\subsection{Non-covariant observables}
So far, we have concentrated on covariant observables $\Ao_\Lambda$ and $\Bo_\Gamma$.
Let us have a short look on a class of non-covariant observables.
Let $p$ and $r$ be two probability distributions on $\Z_d$.
We define
\begin{equation*}
\Ao_{\lambda;p}(j) := \lambda \Ao(j) + (1-\lambda) p(j) \id
\end{equation*}
and
\begin{equation*}
\Bo_{\gamma;r}(k) := \gamma \Bo(k) + (1-\gamma) r(k) \id \, .
\end{equation*}
It is straightforward to verify that $\Ao_{\lambda;p}$ is $U$-covariant and $V$-invariant iff $p$ is the uniform distribution on $\Z_d$, in which case $\Ao_{\lambda;p}=\Ao_{\lambda}$.
(Analogous statement holds for $\Bo_{\gamma;r}$).
The following result is a generalization of Proposition 5 in \cite{BuHe08}.
\begin{proposition}\label{prop:non}
If $\Ao_{\lambda;p}$ and $\Bo_{\gamma;r}$ are jointly measurable, then $\Ao_{\lambda}$ and $\Bo_{\gamma}$ are jointly measurable.
\end{proposition}
\begin{proof}
Suppose that $\Ao_{\lambda;p}$ and $\Bo_{\gamma;r}$ are jointly measurable and let $\Co$ be their joint observable. As in the proof of Proposition \ref{prop:if-jm-then-cov}, we define the observable $\widetilde{\Co}$, given by
$$
\widetilde{\Co}(j,k) := \frac{1}{d^2}\sum_{x,y\in\Z_d} U_x^\ast V_y^\ast \Co(j+x,k+y) V_y U_x \, , \qquad j,k\in\Z_d \, .
$$
For each $j$, we have
\begin{eqnarray*}
\sum_{k\in\Z_d} \widetilde{\Co}(j,k) &=& \frac{1}{d^2} \sum_{x,y\in\Z_d} U_x^\ast V_y^\ast \sum_{k\in\Z_d} \Co(j+x,k+y) V_y U_x \\
&=& \frac{1}{d^2} \sum_{x,y\in\Z_d} U_x^\ast V_y^\ast \Ao_{\lambda;p}(j+x) V_y U_x \\
&=& \lambda \Ao(j) + (1-\lambda) \frac{1}{d^2} \sum_{x,y\in\Z_d} p(j+x) \id \\
&=& \Ao_{\lambda}(j) \, .
\end{eqnarray*}
In a similar way we obtain $\sum_{j\in\Z_d} \widetilde{\Co}(j,k) = \Bo_{\gamma}(k)$ for every $k$.
Therefore, $\widetilde{\Co}$ is a joint observable for $\Ao_{\lambda}$ and $\Bo_{\gamma}$.
\end{proof}
As a consequence of Propositions \ref{Prop:gammamax} and \ref{prop:non} we conclude the following necessary criterion for joint measurability.
\begin{corollary}\label{cor:nec}
If two observables $\Ao_{\lambda;p}$ and $\Bo_{\gamma;r}$ are jointly measurable, then
\begin{equation}
\gamma \leq \frac{1}{d} \left[ (d-2)(1-\lambda) + 2 \sqrt{(1-d)\lambda^2 + (d-2)\lambda + 1} \right] \, .
\end{equation}
\end{corollary}
A necessary and sufficient inequality for the joint measurability of $\Ao_{\lambda;p}$ and $\Bo_{\gamma;r}$ must contain also $p$ and $r$ in a form or in another.
It is thus clear that Corollary \ref{cor:nec} does not give a sufficient condition.
A necessary and sufficient condition in the case $d=2$ has been obtained in \cite{StReHe08, BuSc10, YuLiLiOh10}.
We remark that a general necessary condition for the joint measurability of two observables on a finite dimensional system has been presented in \cite{MiIm08}.
A comparison to Proposition \ref{Prop:gammamax} shows that this condition is not sufficient.
We leave it as an open problem to find a necessary and sufficient condition for the joint measurability of $\Ao_{\lambda;p}$ and $\Bo_{\gamma;r}$.
\section{Informational completeness}\label{sec:ic}
We will now study the informational completeness of joint observables of $\Ao_\lambda$ and $\Bo_\gamma$.
Let us first recall that the informational completeness of a covariant phase space observable $\Co_T$ is equivalent to the criterion
\begin{equation}\label{eq:ic}
\tr{TU_x V_y} \neq 0 \quad \forall x,y\in\Z_d \, .
\end{equation}
This result has been discussed e.g. in \cite{AlPr77JMP,CaDeLaLe00JMP,DaPeSa04}.
For completeness, we provide a proof in Appendix.
\begin{proposition}\label{prop:icmax}
Suppose $\Ao_\lambda$ and $\Bo_\gamma$ are two observables with $\lambda \notin \{0,1\}$ and $\gamma =\gamma_{\max}(\lambda)$.
Then $\Ao_\lambda$ and $\Bo_\gamma$ have a unique joint observable $\Co$.
The observable $\Co$ is informationally complete if and only if $d$ is odd.
\end{proposition}
\begin{proof}
From Proposition \ref{Prop:gammamax} we know $\Ao_\lambda$ and $\Bo_\gamma$ have a unique joint observable $\Co_T$, generated by the state $T = \kb{\chi_\lambda}{\chi_\lambda}$, with $\chi_\lambda = \alpha_\lambda \varphi_0 + \beta_\lambda \psi_0$.
The informational completeness of $\Co_T$ is equivalent to the condition \eqref{eq:ic}, and
a straightforward calculation gives
\begin{equation*}
\tr{TU_x V_y} = \ip{\chi_\lambda}{U_xV_y\chi_\lambda}= \alpha_\lambda^2 \delta_{x,0} + \beta_\lambda^2 \delta_{y,0} + \frac{\alpha_\lambda \beta_\lambda}{\sqrt{d}} \left( \omega^{-xy} + 1 \right) \, .
\end{equation*}
Let us first notice that $\alpha_\lambda > 0$ and $\beta_\lambda > 0$ since both $\lambda$ and $\gamma$ are nonzero (see \eqref{eq:ab}).
Hence, $\tr{TU_x V_y}=0$ exactly when $\omega^{-xy}=-1$.
The latter condition is equivalent to $2xy\equiv d \mod 2d$.
We conclude that the informational completeness of $\Co_T$ is equivalent to the fact that the equation $2x = d \mod 2d$ has no solution $x\in\Z_d$, and this holds if and only if $d$ is odd.
\end{proof}
In Proposition \ref{prop:icmax} the crucial assumption is that $\gamma =\gamma_{\max}(\lambda)$.
This guarantees that $\Ao_\lambda$ and $\Bo_\gamma$ have a unique joint observable.
If we have $0<\gamma < \gamma_{\max}(\lambda)$, then $\Ao_\lambda$ and $\Bo_\gamma$ have infinitely many joint observables.
In this case it is always possible to choose an informationally complete joint observable, as we prove in the following.
\begin{proposition}\label{prop:icnonmax1}
Suppose $\Ao_\lambda$ and $\Bo_\gamma$ are two observables with $\lambda \notin \{0,1\}$ and $0<\gamma <\gamma_{\max}(\lambda)$.
Then they have an informationally complete covariant joint observable.
\end{proposition}
\begin{proof}
Let $(\gamma_0 , \lambda_0)$ be the intersection of the half line $\R_+ (\gamma , \lambda)$ with the boundary of the domain \eqref{eq:region} in $\R_+^2$, and let $t_0>1$ such that $t_0 (\gamma , \lambda) = (\gamma_0 , \lambda_0)$. Let $\tau = 1 - 1/t_0 \in (0,1)$. We treat separately the cases of odd and even $d$.
1) Suppose that $d=2n$ is even.
For all $k\in\Z_d$, we denote
\begin{equation*}
X_k := \frac{i}{2} \left( \kb{\varphi_{-k}}{\varphi_0} - \kb{\varphi_0}{\varphi_{-k}} + \kb{\varphi_k}{\varphi_0} - \kb{\varphi_0}{\varphi_k} \right) \, .
\end{equation*}
The linear maps $X_k$ are selfadjoint trace zero operators for every $k$, and it is easy to check that
$$
\sum_{x\in\Z_d} U_x V_y X_k V_y^* U_x^* = 0 \, , \qquad \sum_{y\in\Z_d} U_x V_y X_k V_y^* U_x^* = 0 \, .
$$
We introduce the selfadjoint operators
$$
X = \sum_{k\in\Z_d} X_k \, ,
$$
and, for $\kappa > 0$,
$$
S_\kappa = \frac{1}{d} \id + \kappa X \, .
$$
If $\kappa < 1/(d\no{X})$, then $S_\kappa\in\sh$. Moreover, the associated covariant phase space observable $\Co_{S_{\kappa}}$ has trivial marginals $\Ao_0$ and $\Bo_0$.
A straightforward calculation gives
$$
\tr{S_\kappa U_x V_y} = \delta_{x,0}\delta_{y,0} + i \kappa (\omega^{-xy} - 1) \, .
$$
The covariant phase space observable $\Co_{T_\kappa}$ associated to the state $T_\kappa = (1-\tau) \kb{\chi_{\lambda_0}}{\chi_{\lambda_0}} + \tau S_\kappa$, with $\chi_{\lambda_0} = \alpha_{\lambda_0} \varphi_0 + \beta_{\lambda_0} \psi_0$, has margins $\Ao_\lambda$ and $\Bo_\gamma$. Moreover,
\begin{eqnarray}\label{eq:infocom}
\tr{T_\kappa U_x V_y} & = & (1-\tau) \left( \alpha_{\lambda_0}^2 \delta_{x,0} + \beta_{\lambda_0}^2 \delta_{y,0} \right) + \tau \delta_{x,0} \delta_{y,0} \notag\\
&& + (1-\tau) \frac{\alpha_{\lambda_0} \beta_{\lambda_0}}{\sqrt{d}} (\omega^{-xy} + 1) + i\kappa \tau (\omega^{-xy} - 1) \, .
\end{eqnarray}
Let
$$
\varepsilon = \min_{\{k\in\Z_d \, , \, k\neq n\}} |\omega^k + 1| \, , \qquad \delta = \max_{\{k\in\Z_d\}} |\omega^k + 1| \, .
$$
For $\kappa < \min\left\{ \alpha_{\lambda_0} \beta_{\lambda_0} (1-\tau) \varepsilon / (\tau \delta\sqrt{d}) \, , \, 1/(d\no{X}) \right\}$, the right hand side of \eqref{eq:infocom} is nonzero for all $x,y\in\Z_d$, which proves informational completeness of $\Co_{T_\kappa}$ by the criterion \eqref{eq:ic}.
2) Suppose that $d$ is odd. Then, for $T = (1-\tau) \kb{\chi_{\lambda_0}}{\chi_{\lambda_0}} + (\tau/d) \id$, the associated covariant phase space observable $\Co_T$ has margins $\Ao_\lambda$ and $\Bo_\gamma$, and
$$
\tr{TU_x V_y} = (1-\tau) \left[ \alpha_{\lambda_0}^2 \delta_{x,0} + \beta_{\lambda_0}^2 \delta_{y,0} + \frac{\alpha_{\lambda_0} \beta_{\lambda_0}}{\sqrt{d}} \left( \omega^{-xy} + 1 \right) \right] + \tau \delta_{x,0} \delta_{y,0} \, ,
$$
which is nonzero for all $x,y\in \Z_d$.
The informational completeness of $\Co_T$ then follows from the criterion \eqref{eq:ic}.
\end{proof}
The two trivial cases $\lambda=0$ or $\gamma=0$ are not very interesting, but for completeness we make the following observation.
\begin{proposition}\label{prop:icnonmax2}
Suppose $\Ao_\lambda$ and $\Bo_\gamma$ are two observables with $\lambda = 0$ or $\gamma=0$. Then they have no informationally complete joint observable.
\end{proposition}
\begin{proof}
We consider only the case $\lambda=0$, the case $\gamma=0$ being similar.
Suppose that $\Co$ is a joint observable of $\Ao_0 = \mu \id$ and $\Bo_\lambda$.
We have
$$
\sum_{k\in\Z_d} \Co(j,k) = \Ao_0 (j) = \frac{1}{d} \id \qquad \forall j\in\Z_d \, ,
$$
hence
$$
{\rm span}\, \left\{ \Co(j,k) \mid j,k\in\Z_d \right\} = {\rm span}\, \left\{ \id \, , \, \Co(j,k) \mid j\in\Z_d \, , \, k\in\Z_d \setminus \{0\} \right\}
$$
and then, for $d\geq 2$,
$$
\dim {\rm span}\, \left\{ \Co(j,k) \mid j,k\in\Z_d \right\} \leq 1 + d(d-1) < d^2 \, .
$$
Thus, $\Co$ is not informationally complete.
\end{proof}
\section{Sequential implementation of joint observables}\label{sec:seq}
In this section we discuss the sequential implementation of joint observables of $\Ao_\Lambda$ and $\Bo_\Gamma$ in the light of the recent results obtained in \cite{CaHeTo11} and \cite{HeWo10}.
For illustrative purposes, we point out that two naive methods do not work.
\subsection{Nondisturbing measurement}
Suppose that $\Ao_\Lambda$ and $\Bo_\Gamma$ are jointly measurable, i.e., they satisfy the condition stated in Proposition \ref{prop:joint->phi}.
The most uncomplicated way to realize their joint measurement would be to perform an $\Ao_\Lambda$-measurement without disturbing the subsequent $\Bo_\Gamma$-measurement.
In terms of instruments, this would mean that we choose an $\Ao_\Lambda$-compatible instrument $\mathcal{I}$ such that
\begin{equation*}
\sum_j \mathcal{I}_j^\ast(\Bo_\Gamma(k)) = \Bo_\Gamma(k)
\end{equation*}
for all $k$.
However, this type of measurement is typically not possible since a quantum measurement necessarily disturbs the input state.
Let us first notice that $\Ao_\lambda$ and $\Bo_\gamma$ commute if and only if $\lambda\gamma=0$, meaning that one of them is a trivial observable.
Generally, a non-disturbing measurement can be possible even if two observables do not commute.
But applying Proposition 3 from \cite{HeWo10} we see that this possibility is excluded whenever $\Bo_\gamma$ is informationally equivalent with $\Bo$ in the sense that the linear spans of the sets $\{\Bo_\gamma(k):k\in\Z_d\}$ and $\{\Bo(k):k\in\Z_d\}$ are equal.
This property is satisfied by any $\Bo_\gamma$ with $\gamma\neq 0$.
Therefore, whenever both observables are nontrivial, then $\Ao_\lambda$-measurement disturbs the subsequent $\Bo_\gamma$-measurement and the resulting observable is not a joint measurement of $\Ao_\lambda$ and $\Bo_\gamma$.
\subsection{Measuring only part of the ensemble}
Suppose we have a measurement setup for $\Ao$ and that the corresponding instrument is $\mathcal{I}$.
We can implement an unsharp observable $\Ao_\lambda$ by performing the $\Ao$-measurement in a randomly chosen $\lambda$-part of the ensemble and doing nothing for the rest $(1-\lambda)$-part.
The corresponding $\Ao_\lambda$-compatible instrument $\mathcal{I}'$ is then
\begin{equation}\label{eq:doing-nothing}
\mathcal{I}'_j(\varrho) = \lambda \mathcal{I}_j(\varrho) + \frac{1-\lambda}{d} \, \varrho \, .
\end{equation}
This is clearly a very direct way to decrease the disturbance that an $\Ao$-measurement would cause.
By measuring the observable $\Bo$ after the first measurement, one could expect to have a useful joint measurement of $\Ao_\lambda$ and some approximate version of $\Bo$.
However, this type of method does not yield an informationally complete joint measurement.
The observable $\Ao$ consists of rank-1 operators, and any $\Ao$-compatible instrument $\mathcal{I}$ is of the form
\begin{equation}
\mathcal{I}_j(\varrho)= \tr{\varrho \Ao(j)} \xi_j
\end{equation}
for some set of states $\{\xi_j: j\in\Z_d\}$ \cite{HeWo10}.
If we insert this form into \eqref{eq:doing-nothing}, we see that a sequential measurement consisting of $\mathcal{I}'$ followed by a $\Bo$-measurement leads to the joint observable
\begin{equation*}
\Co(j,k) = \lambda\ \tr{\xi_j \Bo(k)} \Ao(j) + \frac{1-\lambda}{d} \, \Bo(k) \, , \qquad j,k\in\Z_d \, .
\end{equation*}
The linear span of the set $\{\Co(j,k):j,k\in\Z_d \}$ is contained in the linear span of the union $\{\Ao(j): j\in\Z_d\}\cup\{\Bo(k): k\in\Z_d\}$.
The latter is strictly smaller than $\lh$, hence $\Co$ is not informationally complete.
We also see that this kind of approach cannot give more information than separate measurements of $\Ao$ and $\Bo$ would give.
\subsection{General joint observables}
We recall from \cite{HeWo10} that every joint observable of $\Ao_\Lambda$ and $\Bo_\Gamma$ can be implemented as a sequential measurement of $\Ao_\Lambda$ followed by a measurement of $\Bo$.
Namely, suppose that $\Co$ is a joint observable of $\Ao_\Lambda$ and $\Bo_\Gamma$.
We define an instrument $\mathcal{I}$ by
\begin{equation}\label{eq:instrument-general}
\mathcal{I}_j(\varrho) = \sum_{k\in\Z_d} \tr{\varrho \Co(j,k)} \Bo(k) \, .
\end{equation}
This is an $\Ao_\Lambda$-compatible instrument, and from $\Bo(k)\Bo(k')=\delta_{k,k'}\Bo(k)$ it follows that
\begin{equation*}
\tr{\Bo(k)\mathcal{I}_j(\varrho)} = \tr{\varrho \Co(j,k)} \, .
\end{equation*}
Hence, $\Co(j,k) = \mathcal{I}_j^\ast (\Bo(k))$, and we conclude that $\Co$ is implemented as a sequential measurement of $\Ao_\Lambda$ followed by a measurement of $\Bo$, as claimed.
\subsection{Covariant phase space observables}
The instrument defined in \eqref{eq:instrument-general} may look quite artificial and before we know the structure of $\Co$, the formula does not give us any hint on the structure of $\mathcal{I}$.
In contrast, every covariant phase space observable can implemented as sequential measurement of $\Ao_\Lambda$ and $\Bo$ in a very specific form.
As explained in \cite{CaHeTo11}, every covariant $\Ao_\Lambda$-compatible instrument gives rise to a covariant phase space observable.
Covariance of an instrument $\mathcal{I}$ here means that
\begin{equation}\label{eq:ins-cov}
U_xV_y\mathcal{I}_j(V_y^\ast U_x^\ast \varrho U_x V_y)V_y^\ast U_x^\ast = \mathcal{I}_{j+x}(\varrho)
\end{equation}
for all $x,y,j\in\Z_d$ and $\varrho\in\sh$.
It is straightforward to verify that the joint observable $\Co(j,k) := \mathcal{I}_j^\ast (\Bo(k))$ is a covariant phase space observable.
We demonstrate this method by choosing the $\Ao_\lambda$-compatible L\"uders instrument $\mathcal{I}^L$, defined as
\begin{equation*}
\mathcal{I}^L_j(\varrho) = \sqrt{\Ao_\lambda(j)} \varrho \sqrt{\Ao_\lambda(j)} \, .
\end{equation*}
It is straightforward to see that $\mathcal{I}^L$ satisfies \eqref{eq:ins-cov}.
The covariant joint observable is then
\begin{equation*}
\Co(j,k) = \sqrt{\Ao_\lambda(j)} \Bo(k) \sqrt{\Ao_\lambda(j)} \, ,
\end{equation*}
and its associated state is
$$
T = d \Co(0,0) = d \sqrt{\Ao_\lambda(0)} \kb{\psi_0}{\psi_0} \sqrt{\Ao_\lambda(0)} \, .
$$
Since
\begin{eqnarray*}
\sqrt{\Ao_\lambda(j)} & = & \sqrt{\frac{1-\lambda}{d}} \, \id + \left( \sqrt{\frac{(d-1) \lambda + 1}{d}} - \sqrt{\frac{1-\lambda}{d}} \right) \, \Ao(j) \\
& = & \frac{\beta_\lambda}{\sqrt{d}} \, \id + \alpha_\lambda \kb{\varphi_j}{\varphi_j}
\end{eqnarray*}
and
$$
\sqrt{\Ao_\lambda(0)} \psi_0 = \frac{1}{\sqrt{d}} \left(\beta_\lambda \psi_0 + \alpha_\lambda \varphi_0\right) = \frac{1}{\sqrt{d}} \chi_\lambda \, ,
$$
we see that $T=\kb{\chi_\lambda}{\chi_\lambda}$, hence by Proposition \ref{Prop:gammamax} the marginal $\Bo_\gamma$ is such that $\gamma=\gamma_{\max}(\lambda)$.
In conclusion, this type of sequential measurement of $\Ao_\lambda$ and $\Bo$ is effectively a joint measurement of $\Ao_\lambda$ and $\Bo_\gamma$ with minimal unsharpnesses.
\section{Discussion}\label{sec:discussion}
In our investigation we have concentrated on canonically conjugated pairs of observables, i.e., the orthonormal bases $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$ have been assumed to be Fourier connected with respect to the Fourier transform of the cyclic group $\Z_d$; see \eqref{eq:defF}. Equivalently, we have assumed that the two bases satisfy $\ip{\varphi_j}{\psi_k} = (1/\sqrt{d}) \, \omega^{jk}$ for all $j,k\in\Z_d$.
As a consequence the observables $\Ao$ and $\Bo$, both defined on $\Omega_\Ao = \Omega_\Bo = \Z_d$ as $\Ao(j)=\kb{\varphi_j}{\varphi_j}$ and $\Bo(k)=\kb{\psi_k}{\psi_k}$, satisfy the covariance and invariance conditions \eqref{eq:cov-A} -- \eqref{eq:cov-B}, which turn out to be very useful in our calculations.
Our approach covers more cases than it may seem at the first sight.
Namely, we recall that two orthonormal bases $\{\varphi_j\}_j$ and $\{\varphi'_j\}_j$ define the same observable iff there are complex numbers $\alpha_j$ with $|\alpha_j|=1$ such that $\varphi'_j=\alpha_j \varphi_j$.
To illustrate an application of this many-to-one correspondence, suppose that the dimension $d$ is an odd prime number, say $d=p$ (the generalization to the case $d=p^r$, with $r$ positive integer, is straightforward).
In this case it is easy to give a full set of $p+1$ MUBs \cite{WoFi89}; fix an orthonormal basis $\{\varphi_j\}_{j\in\Z_p}$ and define $p$ orthonormal bases $\{\psi^a_k\}_{k\in\Z_p}$, each one labeled by $a\in\Z_p$, by
$$
\psi^a_k = \frac{1}{\sqrt{p}} \sum_{x\in\Z_p} \omega^{ax^2 + kx} \varphi_x \, .
$$
The fact that these are MUBs follows from the Gauss summation formula
\begin{equation}\label{eq:Gauss}
\frac{1}{\sqrt{p}} \sum_{x\in\Z_p} \omega^{ax^2} = \left( \frac{a}{p} \right) \times \left\{ \begin{array}{ccc} 1 & \mbox{if} & p \in 4 \nat + 1 \\ i & \mbox{if} & p \in 4 \nat - 1 \end{array} \right. \, ,
\end{equation}
where $\left( \frac{a}{p} \right)$ is the Legendre symbol (see e.g. \cite{BeEv81}).
It is immediate to see that the orthonormal basis $\{\psi^a_k\}_{k\in\Z_p}$ is Fourier connected to the orthonormal basis $\{\varphi'_j\}_{j\in\Z_p}$ given by
$$
\varphi'_j = \omega^{aj^2} \varphi_j \quad \forall j \, ,
$$
i.e., $\ip{\varphi'_j}{\psi^a_k} = (1/\sqrt{p}) \, \omega^{jk}$.
Moreover, for $a,b\in\Z_p \setminus \{0\}$, with $a\neq b$, define the rescaled orthonormal bases $\{\psi^{a\prime}_j\}_{j\in\Z_p}$ and $\{\psi^{b\prime}_k\}_{k\in\Z_p}$, given by
\begin{align*}
\psi^{a\prime}_j & = \omega^{-4^{-1} j^2 (b-a)^{-1}} \psi^a_j \\
\psi^{b\prime}_k & = \omega^{k^2(b-a)} \left( \frac{b-a}{p} \right) \psi^b_{2k(b-a)} \times \left\{ \begin{array}{ccc} 1 & \mbox{if} & p \in 4 \nat + 1 \\ - i & \mbox{if} & p \in 4 \nat - 1 \end{array} \right.
\end{align*}
(Here $\omega^{x^{-1}}$ means `$\omega$ to the inverse of $x$ in the field $\Z_p$' and should not be confused with $e^{\frac{2\pi i}{px}}$). Then, an easy computation using the Gauss formula \eqref{eq:Gauss} yelds
$$
\ip{\psi^{b\prime}_h}{\psi^{a\prime}_k} = \frac{1}{\sqrt{p}} \omega^{-hk} \, ,
$$
which shows that also $\{\psi^{a\prime}_j\}_{j\in\Z_p}$ and $\{\psi^{b\prime}_k\}_{k\in\Z_p}$ are Fourier connected.
More generally, one can start from a complementary pair of observables, which means that $\{\varphi_j\}_j$ and $\{\psi_k\}_k$ are mutually unbiased but not necessarily Fourier connected.
Obviously, we can still ask similar questions on joint measurements.
Especially, it would be interesting to know whether Proposition \ref{prop:symmetric} is still valid under this more general setting.
In other words, the question is whether all complementary pairs are essentially similar with respect to joint measurability
Even if we leave this question open in the general case, we can see that our approach generalizes to a larger domain than we have explicitly used it for.
Indeed, all our results are still valid (and with only very slight modifications in some of the proofs) if we consider Fourier transform with respect to a generic abelian group $G$ with order $d$, i.e., $G=\Z_{d_1}\times\ldots\times\Z_{d_k}$ for $d_1+\cdots+d_n=d$ and $d_i = p_i^{r_i}$, with $p_i$ prime and $r_i$ integer for all $i=1,\ldots ,n$.
In this case, $\hh = \hh_1 \otimes \ldots \otimes \hh_n$ with $\dim \hh_i = d_i$, a basis $\{\varphi^i_j\}_{j\in\Z_{d_i}}$ is chosen in each factor Hilbert space $\hh_i$, and the $G$-Fourier transform of $\hh$ is just the tensor product $\mathcal{F} = \mathcal{F}_1 \otimes \ldots \otimes \mathcal{F}_n$, where each $\mathcal{F}_i$ is the $\Z_{d_i}$-Fourier transform in $\hh_i$ with respect to the basis $\{\varphi^i_j\}_{j\in\Z_{d_i}}$, as defined in \eqref{eq:defF}.
The mutually unbiased bases $\{\varphi_j\}_{j\in\Z_d}$ and $\{\psi_k\}_{k\in\Z_d}$ are replaced by the bases $\{\varphi_{j_1 \, , \ldots,\, j_n}\}_{j_1\in\Z_{d_1} \, , \ldots , \, j_n\in\Z_{d_n}}$ and $\{\psi_{k_1 \, , \ldots,\, k_n}\}_{k_1\in\Z_{d_1} \, , \ldots ,\, k_n\in\Z_{d_n}}$ of $\hh$, given by
\begin{align*}
\varphi_{j_1 \, , \ldots ,\,j_n} & = \varphi^1_{j_1} \otimes \ldots \otimes \varphi^n_{j_n} \\
\psi_{k_1 \, , \ldots ,\,k_n} & = \mathcal{F}^\ast (\varphi_{j_1 \, , \ldots j_n}) = \psi^1_{k_1} \otimes \ldots \otimes \psi^n_{k_n} \, .
\end{align*}
Their associated complementary observables $\Ao$ and $\Bo$ are now both defined on $G$, and given by $\Ao (j_1 \, , \ldots , \, j_n) = \Ao(j_1) \otimes \ldots \otimes \Ao(j_n)$ and $\Bo (k_1 \, , \ldots , k_n) = \Bo(k_1) \otimes \ldots \otimes \Bo(k_n)$. They still satisfy the analogues of the covariance and invariance conditions \eqref{eq:cov-A} -- \eqref{eq:cov-B}, if the representations $U$ and $V$ are replaced by suitable tensor products.
To demonstrate that we can now handle larger class of complementary observables, let $\hi=\C^4$, choose an orthonormal basis $\{ \varphi_j \}_{j\in\{0,\ldots, 3\}}$ of $\C^4$ and set
\begin{align*}
&\psi_0 = \half ( \varphi_0 + \varphi_1 + \varphi_2 + \varphi_3) \, , & \psi_1 = \half ( \varphi_0 - \varphi_1 + \varphi_2 - \varphi_3) \, ,\\
&\psi_2= \half ( \varphi_0 + \varphi_1 - \varphi_2 - \varphi_3)\, ,
&\psi_3 = \half ( \varphi_0 - \varphi_1 - \varphi_2 + \varphi_3) \, .
\end{align*}
Then the observables $\Ao(j)=\kb{\varphi_j}{\varphi_j}$ and $\Bo(k)=\kb{\psi_k}{\psi_k}$ are complementary.
They are equally defined by any two orthonormal bases $\{ \alpha_j\varphi_j \}_{j\in\{0,\ldots, 3\}}$ and $\{\beta_k\psi_k\}_{k\in\{0,\ldots, 3\}}$, where $\alpha_j,\beta_k$ are complex numbers with unit modulus.
If some pair of these orthonormal bases were connected by the $\Z_4$-Fourier transform, then the matrix of their scalar products $[\overline{\alpha_j} \beta_k \ip{\varphi_j}{\psi_k}]$ should be equal to the $\Z_4$-Fourier matrix
\begin{align*}
\frac{1}{2} \begin{pmatrix} 1&1&1&1\\ 1&i&-1&-i \\ 1&-1&1&-1\\ 1&-i&-1&i \\
\end{pmatrix}
\end{align*}
or to a matrix obtained from the above by some permutations of its rows and columns. It is straightforward to verify that the deriving set of equations for $\alpha_j$ and $\beta_k$ has no solution.
However, the matrix of scalar products $[\ip{\varphi_j}{\psi_k}]$ is just the Fourier matrix of $G=\Z_2 \times \Z_2$, i.e.,
\begin{equation}\label{eq:Fmatrix}
\frac{1}{2} \left( \begin{array}{cccc} 1 & 1 & 1 & 1 \\ 1 & -1 & 1 & -1 \\ 1 & 1 & -1 & -1 \\ 1 & -1 & -1 & 1 \end{array} \right) \, .
\end{equation}
In other words, the two orthonormal bases are connected by the Fourier transform of $\Z_2 \times \Z_2$.
\section*{Appendix: criterion for informational completeness}
\begin{theorem}\label{th:infocom}
Let $\Co_T$ be a covariant phase space observable.
Then $\Co_T$ is informationally complete if and only if
\begin{equation}\label{eq:icbis}
\tr{TU_x V_y} \neq 0 \quad \forall x,y\in\Z_d \, .
\end{equation}
\end{theorem}
Our proof of the above theorem relies on the following well known reconstruction formula for the Weyl-Heisenberg group, which is just a special case of orthogonality relations for irreducible representations of compact groups.
\begin{proposition}\label{Prop:recons}
The following reconstruction formula holds for every $A\in\lh$:
\begin{equation}\label{eq:recons}
\frac{1}{d} \sum_{x,y\in\Z_d} \tr{AV^*_y U^*_x} U_x V_y = A \, .
\end{equation}
\end{proposition}
\begin{proof}
For all $h,k\in\Z_d$, we have $\ip{\varphi_k}{U_x V_y \varphi_h} = \omega^{hy} \, \delta_{h+x,k}$.
Thus, for all $h,k,m,n\in\Z_d$, we obtain
\begin{eqnarray*}
&& \ip{\varphi_k}{\left[\sum_{x,y\in\Z_d} \tr{\kb{\varphi_m}{\varphi_n} V_y^\ast U_x^\ast} U_x V_y \right] \varphi_h} = \\
&& \qquad \qquad \qquad \qquad = \sum_{x,y\in\Z_d} \ip{U_x V_y \varphi_n}{\varphi_m} \ip{\varphi_k}{U_x V_y \varphi_h} \\
&& \qquad \qquad \qquad \qquad = \sum_{x,y\in\Z_d} \omega^{(h-n)y} \, \delta_{h+x,k} \, \delta_{n+x,m} \\
&& \qquad \qquad \qquad \qquad = \sum_{y\in\Z_d} \omega^{(h-n)y} \, \delta_{h-n,k-m} \\
&& \qquad \qquad \qquad \qquad = d \, \delta_{h,n} \, \delta_{h-n,k-m} \\
&& \qquad \qquad \qquad \qquad = d \, \delta_{h,n} \, \delta_{k,m} \\
&& \qquad \qquad \qquad \qquad = d \ip{\varphi_k}{\varphi_m} \ip{\varphi_n}{\varphi_h} \, ,
\end{eqnarray*}
which proves \eqref{eq:recons} for $A=\kb{\varphi_m}{\varphi_n}$.
Since every $A\in\lh$ is a linear combination of this type of operators, the claim follows.
\end{proof}
\begin{proof}[Proof of Theorem \ref{th:infocom}]
Let $\ell(\Z_d^2)$ be the linear space of complex functions on $\Z_d^2\equiv\Z_d\times\Z_d$.
We recall that, by Proposition 5.1 in \cite{BuCaLa95}, $\Co_T$ is informationally complete if and only if the linear map
$$
V_T : \lh \to \ell (\Z^2_d) \, , \qquad [V_T (A)] (x,y) = \tr{\Co_T (x,y) A}
$$
is injective.
Since the dimensions of $\lh$ and $\ell (\Z^2_d)$ are both $d^2$, we conclude that $\Co_T$ is informationally complete if and only if $V_T$ is an isomorphism.
We define the following three linear maps
\begin{align*}
\Phi & : \lh\to \hh\otimes\hh \, , \qquad \Phi(A) = \frac{1}{d} \sum_{x,y\in\Z_d} \tr{AV^*_y U^*_x} \, \varphi_x\otimes\varphi_y \\
M_T & : \hh\otimes\hh \to \hh\otimes\hh \, , \qquad M_T (\varphi_x\otimes\varphi_y) = \tr{TU_x V_y} \, \varphi_x\otimes\varphi_y \\
R & : \hh\otimes\hh \to \ell (\Z^2_d) \, , \qquad R \phi (x,y) = \ip{\varphi_y\otimes\varphi_x}{\phi} \, .
\end{align*}
The map $R$ is clearly a linear isomorphism, $\Phi$ is a linear isomorphism by Proposition \ref{Prop:recons}, and $M_T$ is a linear isomorphism if and only if \eqref{eq:icbis} holds.
We now evaluate the composition map $R(\mathcal{F}\otimes\mathcal{F}^* )M_T \Phi$.
For all $A\in\lh$, we obtain
\begin{eqnarray*}
&& [R(\mathcal{F}\otimes\mathcal{F}^* )M_T \Phi (A)](h,k) = \\
&& \qquad = \frac{1}{d} \sum_{x,y\in\Z_d} \tr{TU_x V_y} \tr{AV^*_y U^*_x} \ip{\varphi_k\otimes\varphi_h}{(\mathcal{F}\otimes\mathcal{F}^*)(\varphi_x\otimes\varphi_y)} \\
&& \qquad = \frac{1}{d^2} \sum_{x,y\in\Z_d} \tr{TU_x V_y} \tr{AV^*_y U^*_x} \omega^{yh-xk} \\
&& \qquad = \frac{1}{d^2} \sum_{x',y'\in\Z_d} \omega^{y'h-x'k} \tr{TU^*_h U_{x'} V^*_k V_{y'}} \tr{AV_k V^*_{y'} U_h U^*_{x'}} \\
&& \qquad = \frac{1}{d^2} \sum_{x',y'\in\Z_d} \tr{TU^*_h V^*_k U_{x'} V_{y'}} \tr{AV_k U_h V^*_{y'} U^*_{x'}} \\
&& \qquad = \frac{1}{d} \tr{TU^*_h V^*_k AV_k U_h} \\
&& \qquad = [V_T (A)] (h,k)
\end{eqnarray*}
(in the third equality we set $x=x'-h$, $y=y'-k$, in the fourth we used the commutation relation for $U$ and $V$, and in the fifth we applied the reconstruction formula \eqref{eq:recons}). As the map $R(\mathcal{F}\otimes\mathcal{F}^* )M_T \Phi$ is an isomorphism if and only if \eqref{eq:icbis} holds, the same is true for the map $V_T$, and the theorem follows.
\end{proof}
\section*{Acknowledgements}
T.H. is grateful to Cosmo Lupo and Mario Ziman for illuminating discussions on qubit tomography.
T.H. acknowledges financial support from the Academy of Finland (grant no. 138135) and the Magnus Ehrnrooth foundation.
|
\section{Introduction}
It is well known theoretically
that systems of reduced dimensionality, especially in one
dimension, present simultaneously enhanced
quantum fluctuations and stronger interaction effects that can lead to
exotic ground
states~\cite{varma_nonfermiliquid_review,schulz_houches_revue}.
>From the experimental point of view, there
are many prototypical one-dimensional systems, that range from
organic~\cite{bourbonnais_jerome_review,bourbonnais2008} or
inorganic~\cite{mizokawa2002,wang06_arpes_limo6o17} conductors,
and antiferromagnetic (AF) spin chain~\cite{hammar1999,lake2005} or
ladder~\cite{dagotto_supra_ladder_review,klanjsek08_bpcp} materials,
to nanoscale systems such as
of quantum wires~\cite{auslaender_quantumwire_tunneling,hilke2001},
carbon
nanotubes~\cite{bockrath_luttinger_nanotubes,ishii03_nanotube_pes,gao2004,lee2004}
or self organized Au atomic wires on Ge(001) semiconductor
surfaces~\cite{blumenstein2011}. More
recently, advances in atom trapping technology has permitted the
realization of both fermionic and bosonic one-dimensional systems with
unprecedented control~\cite{paredes_toks_experiment,kinoshita_tonks_experiment,liao2010,cazalilla_review_11}. The low-energy physics of
such one-dimensional systems is well described by the Tomonaga-Luttinger Liquid (TLL) theory
\cite{schulz_houches_revue,varma_nonfermiliquid_review,voit_bosonization_revue,giamarchi_book_1d}. In
a single component TLL, there is a single gapless branch of
excitations with linear dispersion,
and the interplay between interactions and quantum fluctuations in
the ground state leads to
power-law decay of correlations with interaction dependent
exponents. Remarkably, the low energy theory is fully characterized by
two parameters: the velocity $u$ of the linearly dispersion
excitations, and the dimensionless exponent $K$
controlling the decay of all correlations, the corresponding exponents
being rational functions of $K$. In physical systems several prominent
features of TLL have been observed after measuring
the spectral
function~\cite{bourbonnais_jerome_review,wang06_arpes_limo6o17}, the
structure factor~\cite{lake2005} or the conductivity~\cite{bockrath_luttinger_nanotubes}, and more recently the first
quantitative check of TLL physics has appeared for the spin-1/2
ladder material bis(piperidinium)
tetrabromocuprate(II) (C$_5$H$_{12}$N)$_2$CuBr$_4$ (abbreviated BPCB in the following), in an applied magnetic field~\cite{klanjsek08_bpcp}.
However, despite this recent achievement, in many of the physical systems mentioned above, little control can be exerted on the values of $u$ and
$K$ and thus the Luttinger exponent $K$ is taken as an
adjustable parameter~\cite{bourbonnais_jerome_review,bockrath_luttinger_nanotubes}. This fact prompts for the search of more than
one signature of Tomonaga-Luttinger liquid physics for a single
system.
In the case of systems with strong
confinement (e.g. confined quantum gases), excitation properties
can be most easily accessed by light spectroscopy techniques, as
proposed in the early days of atomic Bose-Einstein
Condensation~\cite{Javanainen95,Graham96}. For
example, the spectral function has recently been measured in trapped
Fermi gases by radiofrequency spectroscopy~\cite{one} and the dynamical
structure factor has been studied successfully by optical Bragg
spectroscopy in free and trapped Bose-Einstein condensates~\cite{two,three, four, five, six} as well as trapped Fermi gases~\cite{seven}.
Bragg spectroscopy can be based on energy transfer to the system at
fixed momenta~\cite{eleven,twelve,thirth} or can permit the study of
the full momentum composition of excitations by a coherent momentum
transfer mapping~\cite{ernst_bragg}. For these reasons,
Bragg spectroscopy can be especially useful to investigate
the properties of the many phases
realizable in these systems such as Mott insulator, Tonks-Girardeau
gas or supersolid phases as recently
proposed~\cite{fourth,fiveth,sixth,seventh,eightth,nineth,twentyth,
twentyoneth,golovach_2009}. The most recent experimental progress in producing
long-lived ground-state polar molecules in
a three-dimensional (3D) optical lattice, and possibly also in 1D arrays of
pancakes and 2D arrays of tubes~\cite{Chotia_2011} as well as
condensates of dipolar atoms~\cite{pfau_a0,lu2011}, opens up wide perspectives
in the comprehension of controlled quantum systems with tunable short and long
range interactions under progressively reduced dimensionality.
As we have more extensively reviewed in Ref.~\cite{ours}, among many
possible realizations, quantum dipolar
gases in 1D confinement are quite peculiar TLL systems. Here in fact, one single
parameter drives the crossover from weak to strong interaction
regimes, where however the weakest regime is a
Tonks-Girardeau state, the strongest being a Density Dipolar Wave
state
characterized by quasi-ordering.
Based on the above motivations, we derive an
analytic expression for light-scattering intensity in
the case of a weakly inhomogeneous TLL. This expression is valid within a
Thomas-Fermi description, where the system can be considered locally
homogeneous. The expression requires the knowledge of the density
dependence of the ground-state energy of the homogeneous system, as
can be obtained by {\em e.g.} approximate calculations, exact
Bethe-Ansatz technique or
Quantum Monte Carlo (QMC) simulations. The paper is organized as follows.
After reviewing in
Sec.~\ref{sec:LS_h} the calculation of the dynamic structure
factor and the inelastic light-scattering cross section of homogeneous
Tomonaga-Luttinger Liquids, we derive in Sec.~\ref{sec:LSin_h} the general
expression for the inhomogeneous system within the Thomas-Fermi
approach, in terms of the eigenvalues and
eigenfunctions of the hydrodynamic TLL.
We then specialize in Sec.~\ref{sec:TLLdip} to the case of
one-dimensional quantum bosonic gases with dipolar interaction in an
harmonic trap, using our previous QMC
findings~\cite{ours}. Here the results are explicitly discussed
in the various regime while the single parameter built-up from
density and interaction strength is tuned.
\section{Light-scattering cross section in homogeneous
Tomonaga-Luttinger Liquids }~\label{sec:LS_h}
The dynamic structure factor $S(\boldmath{q},\omega)$ is central
in the description of interacting many-body systems.
$S(\boldmath{q},\omega)$ is related to the Fourier transform of the
imaginary density-density
correlation function with the fluctuation-dissipation theorem.
It is therefore accessible
by means of inelastic scattering, where density fluctuations are
induced in the system and their subsequent relaxation
is measured revealing the system characteristics.
While inelastic neutron
scattering has been the tool to probe the condensate nature of
superfluid helium and the roton spectrum~\cite{Sokol_BEC},
inelastic light scattering has been proposed
and widely used in dilute quantum degenerate
gases. Within linear response theory the scattering
cross section $\sigma$ of light at frequency $\omega$ and
angle $\Omega$ incident on a Bose atomic sample is:
\begin{equation}
\label{eq:cross-section}
\frac{d^2\sigma}{d\Omega d\omega} \propto \frac{1}{\pi
n}(n_B(\omega)+1) \mathrm{Im} \chi(q,\omega)=S(q,\omega)\; ,
\end{equation}
where $n_B(\omega)$ is the Bose distribution function, $n=N/V$ and
$\chi(q,\omega)$ is the Fourier transform of the density-density
correlation function
\begin{equation}
\chi(r,t)=-i \theta(t)\langle[n(r,t),n(0,0)]\rangle.
\end{equation}
Earlier experimental
studies~\cite{three} have shown that condensate properties of atomic cold gases
could be studied by means of Bragg scattering yielding large energy
resolution and sensitivity. The system is illuminated by two lasers
beams of momenta $\boldmath{k}_1$ and $\boldmath{k}_2$ and frequencies $\omega_1, \omega_2$ of difference $\omega$ that creates a periodic field whose intensity is
proportional to $\cos{[(\boldmath{k}_1-\boldmath{k}_2)\cdot
\boldmath{r}-\omega t]}$. The external potential couples to the
density $ n(\boldmath{q})$ of the system where
$\boldmath{q}=\boldmath{k}_1-\boldmath{k}_2$. After using the golden
rule, the response of the
system to this perturbation is the dynamical
structure factor~\cite{brunello2001}. Light scattering experiments then
directly measure $S(\boldmath{q},\omega)$.
This quantity is then a benchmark against the theoretical descriptions
of the systems. For an homogeneous Tomonaga-Luttinger liquid occuring in
interacting one-dimensional system the dynamic structure factor can
be readily obtained~\cite{giamarchi_book_1d}. In the following,
we briefly sketch the derivation.
For a system of interacting spinless particles, either bosons or fermions, the low-energy physics is that of a
Tomonaga-Luttinger liquid whose Hamiltonian is
\begin{equation}
\label{eq:hamiltonian-gases}
H=\int \frac{dx}{2\pi} \left[ u K (\pi \Pi)^2 + \frac u K
(\partial_x \phi)^2 \right],
\end{equation}
with $u$ the velocity of the excitations and $K$ the Tomonaga-Luttinger
exponent. The density operator $n(x)$ is
expressed in terms of bosonic operators $\phi$:
\begin{equation}
\label{eq:bosonized-cold}
n(x)=n_0 -\frac 1 \pi \partial_x \phi + \sum_m A_m \cos (2m(
\phi(x) - \pi n_0 x)),
\end{equation}
with $m$ an integer and $n_0$ the equilibrium density.
If the wavelength of the incoming light is
much larger than the average interparticle distance, we can neglect
the contribution of the oscillatory terms in
Eq.~(\ref{eq:bosonized-cold}). Using translational
invariance, the expression for the density-density response function becomes:
\begin{equation}
\label{eq:boso-commutator}
\chi(x-x',t)= i \frac{\theta(t)}{\pi^2} \langle
[\partial_x \phi(x,t), \partial_x \phi(x',0)]\rangle.
\end{equation}
Knowing that the time-ordered correlation function $\langle T_\tau \lbrack
\phi(x,\tau)-\phi(0,0)\rbrack^2\rangle=K F_1(x,\tau)$ with
$F_1(x,\tau)=
\log \lbrack {(x^2+(u|\tau|+a)^2)/a^2}\rbrack/2$, the imaginary part
of the response function (\ref{eq:boso-commutator})
can be obtained~\cite{giamarchi_book_1d} as
\begin{equation}
\label{eq:ima_chi_ho}
\mathrm{Im} \chi(q,\omega)=\frac{q^2}{2 \omega} u K \left[
\delta(\omega+ u |q|)-\delta(\omega-u|q|)\right]\; \; ,
\end{equation}
giving the scattered intensity at zero temperature:
\begin{eqnarray}
\label{eq:intens-hom}
\frac{d^2\sigma}{d\Omega d\omega} \propto S(q,\omega)&=&
\mathrm{sign}(\omega)\mathrm{Im} \chi(q,\omega)\nonumber \\
&=& \frac{K |q|} 2 [\delta(\omega +
u |q|)+ \delta(\omega - u |q|)].
\end{eqnarray}
Expression (\ref{eq:intens-hom})
embodies the symmetry with respect to inversion of the velocity $u$ as
required by Galilean invariance, and evidences the dependence
of the light-scattering signal from the ratio $q/\omega$.
\section{Light-scattering cross section in inhomogeneous
Tomonaga-Luttinger Liquids}~\label{sec:LSin_h}
\subsection{Hydrodynamic approach}\label{sec:hydro}
The presence of an external potential $V(x)$ confining the
cold atomic cloud induces density inhomogeneity, and the external
light perturbation probing the density-density correlation function
introduces time-dependent processes.
The treatment of the problem is
easier under conditions of weak inhomogeneity and slow processes
as they can be met in experiments, where
external potentials vary on length
and time scales longer than the characteristic system
quantities, and local equilibrium hydrodynamic behavior sets in.
Under these conditions,
the gas can be still described by a hydrodynamic Tomonaga-Luttinger Liquid
Hamiltonian~\cite{maslov_pure_wire,safi_pure_wire,fazio_thermal_1d,petrov_trapped_bosons,petrov04_bec_review,ours}
\begin{equation}
\label{eq:ham-non-uniform}
H_{TLL}=\int^{R}_{-R} \frac{dx}{2 \pi} \left[ u(x)K(x)\pi^2 \Pi(x)^2
+\frac{u(x)}{K(x)} (\partial_x \phi(x))^2 \right]\; .
\end{equation}
Here, the boundary conditions imposed are $\phi(-R)=0$ and
$\phi(R)=-\pi N$, with $N$ the number of particles in the system.
The parameters $u(x)$ and $K(x)$ now depend on position. In analogy with the
homogeneous case, where $u$ and $K$ are related by the expressions
$u/K=(\hbar\pi)^{-1}\partial\mu/\partial n$ and by
Galilean invariance $uK=\pi\hbar n/m$, one sets:
\begin{eqnarray}
\label{eq:uK}
u(x)K(x) &=& \pi \frac{\hbar}{m} n_0(x) \\ \frac{u(x)}{K(x)} &=&
\frac{1}{\hbar \pi} \left( \frac{\partial \mu(n)}{\partial
n}\right)_{n=n_0(x)}
\end{eqnarray}
Once an estimate of the equilibrium density $n_0(x)$
and of the chemical potential $\mu(n)$ are known, this
phenomenological approach allows the determination of $u(x)$ and
$K(x)$.
The response function~(\ref{eq:boso-commutator})
in the case of the Hamiltonian~(\ref{eq:ham-non-uniform}) can be
calculated using the
decomposition:
\begin{eqnarray}
\label{eq:decomposition-eigen}
\nonumber \phi(x)&=& -\pi \frac{\int_{-R}^x dx' \frac
{K(x')}{u(x')}}{\int_{-R}^R dx' \frac{K(x')} {u(x')}} N \\ &+&
\sum_n \sqrt{\frac{\pi}{2\omega_n}} (a^\dagger_n + a_n) \varphi_n(x)\; .
\end{eqnarray}
Here, $[a_n,a^\dagger_n]=\delta_{n,m}$ and the first term comes from the addition of $N$ particles in the system. The functions $\varphi_n$ satisfy the eigenvalue equation:
\begin{eqnarray} \label{eigen}
-\omega_n^2 \varphi_n = u(x) K(x) \partial_x \left(\frac{u(x)}{K(x)}
\partial_x \varphi_n \right)\; ,
\end{eqnarray}
with boundary conditions $\varphi_n (\pm R)=0$,
and the normalization
\begin{eqnarray}
\label{eq:orthogonality}
\int dx \frac{\varphi_n(x) \varphi_m(x)}{u(x) K(x)}= \delta_{n,m}\; .
\end{eqnarray}
The influence of the trapping potential enters eq.~(\ref{eigen})
{\it via} the equations for $u(x)$ and $K(x)$ (\ref{eq:uK}).
The density-density response function thus can be expressed as:
\begin{equation}
\label{eq:eigen-expansion}
\chi(x,x',t)=\theta(t)\sum_n \frac{1}{\pi\omega_n}
\frac{d\varphi_n}{dx} \frac{d\varphi_n}{dx'} \sin(\omega_n t),
\end{equation}
Taking the Fourier transforms with respect to $x$ and $x'$ and the
Laplace transform with respect to $t$, we find:
\begin{eqnarray}
\label{eq:laplace-fourier}
\chi(q,z)=\sum_n \frac{q^2 |\hat{\varphi}_n(q)|^2}{2\pi \omega_n}
\left(\frac{1}{z+\omega_n} -\frac{1}{z-\omega_n}\right),
\end{eqnarray}
where $\mathrm{Im}(z)>0$.
Finally, taking the limit $z \to \omega +i0_+$ we obtain:
\begin{eqnarray}
\nonumber \mathrm{Im} \chi (q,\omega +i0_+)&=& \frac{q^2}{2\omega}
\sum_n |\hat{\varphi}_n(q)|^2 [\delta(\omega-\omega_n) \\ &+&
\delta(\omega+\omega_n)]\; .
\label{eq:ima_chi_noho}
\end{eqnarray}
Eq.~(\ref{eq:ima_chi_noho}) maintains the structure of its homogeneous
counterpart (\ref{eq:ima_chi_ho}).
The density-density response
function can be determined whenever the density dependence of
the ground state energy per unit length $e(n)$ or of the chemical potential
$\mu(n)=(\frac{\partial e}{\partial n})|_{n=n(x)}$
is known. An especially simple situation is realized when
$e(n)\propto n^{\gamma +2}$. That type of dependence of energy
on density corresponds to several limiting cases of 1D TLL systems.
For example, in the Lieb-Liniger gas~\cite{lieb_bosons_1D,lieb_excit}
there are two well understood limits. At low density or strong
repulsion, the gas behaves as a hard-core boson
gas~\cite{girardeau_bosons1d} with $\gamma=1$, while at high density or
weak repulsion, the Bogoliubov approximation applies and gives and
energy density proportional to $n^2$, so that $\gamma=0$.
The study of the crossover between these two limits requires the
Bethe-Ansatz computation of the ground state energy
density~\cite{lieb_bosons_1D}. A similar situation
occurs in the case of dipolar gases. For low densities, the energy per
unit length $e(n)$ has the $\gamma=1$ behavior typical of hard core bosons,
while for high density it has the $\gamma=2$ behavior of a
crystal of classical dipoles, and a Dipolar-Density-Wave
manifests~\cite{ours}. As density increases,
the system crosses over from the low density
hardcore boson gas to the high density Dipolar-Density-Wave.
In the model with $e(n)=g n^{\gamma+2}$ and in the case of harmonic
trapping potential $V(x)=m\Omega_0^2x^2/2$, the
eigenvalues $\omega_n$ of (\ref{eq:decomposition-eigen}) can be found
exactly, and the functions $\varphi_n$ are expressible in terms of
Gegenbauer polynomials~\cite{menotti02_bose_hydro1d,petrov04_bec_review} as:
\begin{eqnarray}
\label{eq:ultraspherical}
\varphi_n(x) &=& A_n \left(1-\frac {x^2}{R^2}\right)^{\alpha+1/2}
C_n^{(\alpha+1)}\left(\frac x R\right), \\ \omega_n^2 &=& \frac
{u_0^2}{R^2} (n+1) (n+2\alpha+1)\; .
\end{eqnarray}
Here, $u_0$ and $K_0$ are the Tomonaga-Luttinger parameters corresponding to
the density at the trap center,
\begin{equation}
A_n=\sqrt{\frac{u_0 K_0}{R}
\frac {n! (n+\alpha+1)}{\pi \Gamma(n+2\alpha+2) }} 2^{\alpha+1/2}
\Gamma(1+\alpha)\; ,
\end{equation}
and $\alpha=(\gamma+1)^{-1} - 1/2$. In particular,
in the case of hard-core Bose gas when $\gamma=1$, $\alpha=0$ and
the Gegenbauer polynomials reduce to Chebyshev
polynomials~\cite{abramowitz_math_functions}. In order to calculate
the scattered light intensity, we need the Fourier transform of the
$\varphi_n$'s. Using Eq.~(7.321) of
Ref.~\onlinecite{gradshteyn80_tables} we obtain:
\begin{eqnarray}
\label{eq:integral-gegenbauer}
|\hat{\varphi}_n(q)|^2 &=& 2 u_0 K_0 R (n+\alpha+1)
\frac{\Gamma(n+2\alpha+2)}{\Gamma(n+1)}\nonumber\\
&& \times \frac{J_{n+\alpha+1}^2(qR)}{(qR)^{2\alpha+2}}\; .
\end{eqnarray}
where the $J_m$ are the Bessel functions of the first kind.
Thus:
\begin{eqnarray} \label{eq:final-boso}
&&\mathrm{Im} \chi(q,\omega+i0_+)= \frac{u_0 K_0}{R \omega} \sum_n
(n+\alpha+1) \frac{\Gamma(n+2\alpha+2)}{\Gamma(n+1)}\nonumber\\
&& \times \frac{J_{n+\alpha+1}^2(qR)}{(qR)^{2\alpha}}
[\delta(\omega-\omega_n)+ \delta(\omega+\omega_n)]
\end{eqnarray}
Eq. (\ref{eq:final-boso}) shows the main features of the scattered
light intensity. This is a set of discrete peaks, whose weight is a function
of $qR$, and whose spacing reduces with increasing the trap size
$R\to \infty$.
\subsection{Approach {\it via}
Density-Functional Theory with Local Density Approximation}
\label{sec:approx}
In the present section we derive an approximate expression for the
dynamical structure factor of an inhomogeneous 1D TLL,
reverting to the Density Functional Theory (DFT) accompanied by a Local Density
Approximation (LDA). We sketch in the following the main concepts and
derivation.
Through the Hohenberg and Kohn theorem, DFT
establishes that
the ground state energy of a system subjected to an external potential
$V(x)$ is a functional
$E_g[n(x)]=E[n(x)]+ \int_{-\infty}^{+\infty} n(x) V(x)dx$
of the density $n(x)$, where $E[n(x)]$
embodies the kinetic and exchange-correlation parts.
The equilibrium density profile is determined by the variational
condition
\begin{eqnarray}
\label{eq:DFT}
\frac{\delta E_g[n(x)]}{\delta n(x)}=\mu\; ,
\end{eqnarray}
stating that equilibrium corresponds to a minimum of the energy against
changes in the particle density, while the total number of particles
is fixed through the (density-dependent)
chemical potential $\mu$. Eq. (\ref{eq:DFT})
reminds the Thomas-Fermi equilibrium condition in non-interacting
systems, and in fact the Density Functional sets a one-to-one
correspondence between the ground state energies of an interacting
system and of its non-interacting analogue.
Whenever an analytic expression of $\mu(n)$ is available,
inversion of the equation of state (\ref{eq:DFT}) allows the
determination of the equilibrium density $n_0(x)$.
While Eq. (\ref{eq:DFT}) is exact, the actual determination of
the functional $E[n(x)]$ needs approximations. Under the conditions of
shallow confinement,
we can safely use the Local Density Approximation. Here,
the functional $E[n(x)]$ is replaced by
\begin{equation}
\label{eq:LDAfunc}
E^{LDA}[n(x)]=\int
e^{hom}[n(x)] n(x) dx \; ,
\end{equation}
where $e^{hom}(n)$ is the energy per particle of the
homogeneous system with density $n$.
Differentiating $E^{LDA}[n(x)]=E_g[n(x)]+\int dx (V(x)-\lambda)
n(x)$ with respect to $n(x)$, $\lambda$ being a Lagrange multiplier
fixing the total number of particles, one obtains the condition for
the local chemical potential
\begin{equation}
\label{eq:thomas-fermi_1}
\mu[ n(x)]= V(R)-V(x),
\end{equation}
where the local chemical
potential is defined by the functional derivative:
\begin{eqnarray}
\label{eq:equation_of_state}
\mu(n)=\frac{\delta E}{\delta n(x)}=\left(\frac{\partial (n e^{hom}(n))}
{\partial n}\right)_{n=n_0(x)}\; .
\end{eqnarray}
If an analytic expression of $\mu(n)$ is given,
Eq.(\ref{eq:equation_of_state}) would allow to find $ n(x)$ by
inverting the relation Eq.~(\ref{eq:thomas-fermi_1}).
The energy $e^{hom}(n)$ can be
obtained after perturbation theory, or by exact calculations such as
Bethe-Ansatz, or else by computational Quantum Monte Carlo methods.
We now turn to the problem of determining the dynamical structure factor of the
inhomogeneous system. To this aim,
we follow the reasoning in ~\cite{vignolo2001,golovach_2009} and
imagine to slice it into small segments of length $\Delta x$, where the
density $n_0(x)$ can be considered uniform, and thus sum together
all the contributions~(\ref{eq:intens-hom}) of the different segments.
The dynamical structure factor of the inhomogeneous system
would then be approximated by:
\begin{eqnarray}
\label{eq:approx-tosi}
S(q,\omega)=\int \frac{dx}{2R} S_{\mathrm{hom}}(q,\omega,n_0(x))\; .
\end{eqnarray}
$S_{\mathrm{hom}}(q,\omega,n)$ is given by
Eq.~(\ref{eq:intens-hom}), where now the Tomonaga-Luttinger
parameters $u=u(n)$ and $K=K(n)$ depend on density.
With the help of (\ref{eq:intens-hom}), we obtain:
\begin{eqnarray}
S(q,\omega)&&= \frac {|q|} {4R} \int_{-R}^R dx K(n_0(x))\\
&& [\delta(\omega - u(n_0(x)) |q|) + \delta(\omega + u(n_0(x))
|q|) ]\nonumber
\end{eqnarray}
Introducing $x^*(\omega/|q|)$, such that $\omega = u(n(x^*))|q|$ we
can rewrite:
\begin{eqnarray}
\label{eq:s-lda}
S(q,\omega)=\frac{K(n_0(x^*))}{2R \left|\frac{du}{dn}\right|_{n=n_0(x^*)}
\left| \frac {d n_0}{dx}\right|_{x=x*}}
\end{eqnarray}
Since the compressibility is a positive quantity, the chemical
potential is an increasing function of the density. Moreover
for a trapping potential that is an increasing function
of position, from Eq.~(\ref{eq:thomas-fermi_1})) the density is seen to
decrease with position. Thus, when the velocity is an increasing
function of density, the solution $x^*$ turns out to be unique
The quantity $\frac {d n_0}{dx}$ can be obtained by differentiating the
relation~(\ref{eq:thomas-fermi_1}) with respect to $x$ , i.e.:
\begin{eqnarray}
\left(\frac{d^2 e}{dn^2}\right)_{n=n_0(x)} \frac{d n_0}{dx} +
\frac{dV}{dx}=0\; .
\end{eqnarray}
We can therefore write:
\begin{eqnarray}
\label{eq:Sqw}
S(q,\omega)=\frac{K(n_0(x^*))\left|\frac{d^2e}{dn^2}\right|_{n=n_0(x^*)}}{2R\left|\frac{du}{dn}\right|_{n=n_0(x^*)}
\left| \frac {dV}{dx}\right|_{x=x*}}.
\end{eqnarray}
We now use the relation $u^2(n)=\frac n m \frac{d^2 e}{dn^2}$ obtained
from (\ref{eq:uK}) and rewrite (\ref{eq:Sqw}) as:
\begin{eqnarray}
\label{eq:s-lda-master}
S(q,\omega)=\frac{\pi \hbar}{R \left|\frac {dV}{dx}\right|_{x=x^*}}
\frac{n_0(x^*)}{\left|1+n_0(x^*) \frac
{e'''(n_0(x^*))}{e''(n_0(x^*))}\right|},
\end{eqnarray}
with the notations $e'(x)=de/dn$, $e''(n)=d^2e/dn^2$, and
$e'''(n)=d^3e/dn^3$.
Formula (\ref{eq:s-lda-master}) represents the main result of this
paper. It gives an analytical expression for the light scattering
cross-section of an inhomogeneous TLL once the ground state energy as a
function of the density is known, e.g. by an exact analytical
(Bethe-Ansatz) or via numerical simulations (QMC). Remarkably,
Eq.~(\ref{eq:s-lda-master}) predicts that $S(q,\omega)$ is only a function of
$\omega/|q|$. In fact, this is the specific signature of
Tomonaga-Luttinger Liquid behavior in shallow trapped 1D Bose systems, as it
can be measured by Bragg spectroscopy.
In order to illustrate the relevant features and make the connection
with Eq. (\ref{eq:final-boso}) obtained {\it via}
the hydrodynamic approach of Sec.~\ref{sec:hydro}, we now treat the
case of harmonic trapping. In this case
$dV/dx=m\Omega_0^2 x$, and using Eq.~(\ref{eq:thomas-fermi_1}) we have
$m\Omega_0^2 |x^*|=\sqrt{2 m\Omega_0^2(e'(n_0(0))-e'(\rho^*))}$,
where we have set $\rho^\star=n_0(x^\star)$ and $u(\rho^*)=\omega/|q|$.
Eq. (\ref{eq:s-lda-master}) thus simplifies into:
\begin{eqnarray}
\label{eq:s-lda-harmonic}
S(q,\omega)=\frac{\pi \hbar \rho^*}{R \sqrt{2 m \Omega_0^2
(e'(n_0(0))-e'(\rho^*))}\left|1 + \rho^*
\frac{e'''(\rho^*)}{e''(\rho_*)}\right|}.
\end{eqnarray}
We now check the consistency of the result (\ref{eq:s-lda-master})
with (\ref{eq:final-boso}), by explicitly calculating (\ref{eq:s-lda-harmonic})
for the model $e(n)\propto n^{\gamma
+2}$. Eq.~(\ref{eq:s-lda-harmonic}) then reads:
\begin{eqnarray}\label{eq:power-case}
S(q,\omega)=\frac{\pi \hbar}{(\gamma+1)m\Omega_0^2 R^2} \left(\frac{m
\Omega_0^2 R^2}{2g(\gamma+2)}\right)^{\frac 1 {\gamma+1}}
\frac{\left(\frac{\omega}{u_0 q}\right)^{\frac 2
{\gamma+1}}}{\sqrt{1-\left(\frac{\omega}{u_0 q}\right)^2}},
\end{eqnarray}
where we have defined $u_0=u(n_0(0))$ as the velocity of excitations in
a uniform system having a density equal to that at the trap center.
We first notice that the dynamical structure factor in
(\ref{eq:power-case}) makes explicit the characteristic already
embodied in the structure of Eq. (\ref{eq:s-lda-harmonic}), namely
that $S(q,\omega)$ depends on wavevector and frequency
solely through their ratio
$\omega/q$. Second, the formula (\ref{eq:power-case})
with $\gamma=1$ agrees with the result of
Ref.~\cite{vignolo2001}, in the limiting $\omega \gg q^2/2$
case. Finally, in
App. \ref{sec:sums} we show by inspection
that the LDA approximation (\ref{eq:power-case}) is fully recovered
from expression (\ref{eq:final-boso}).
\begin{figure}[t]
\centering \includegraphics[scale=0.2]{fig1.ps}
\caption{(Color online).
TLL model with $e(nr_0)\propto n^\gamma$ and $\gamma=2$
in a harmonic trap.
$S(q,\omega)$ in arbitrary units in the ($\omega, q$) plane
and different densities at the trap center.}
\label{fig:bragg3d-plot}
\end{figure}
Fig.~\ref{fig:bragg3d-plot} displays the 3D plot of $S(q,\omega)$
resulting from the use of
(\ref{eq:power-case}) in the $(\omega,q)$ plane, while varying the
densities at the trap center. $S(q,\omega)$ is a set of discrete peaks
whose position varies linearly with $\omega/q$ and such linear
behavior is independent on the interaction strength.
Before proceeding to apply Eq. (\ref{eq:s-lda-harmonic}) to a dipolar
1D Bose gas, we step on commenting
the found correspondence between
hydrodynamic and DFT-LDA approaches on a more general footing.
It is well known for normal Fermi
systems~\cite{Vignale97} with extension to Bose
superfluids~\cite{Chiofalo98}, that the treatment of dynamical
processes in interacting inhomogeneous systems do
require the development the Current-Density Functional Theory, where
invariance conditions render the energy to be a functional of the
current besides density. It was demonstrated that the analogue of LDA
leads in this case to Navier-Stokes equations
(Landau-Khalatnikov two-fluid equations for superfluids),
where viscosities, densities and currents (normal and superfluid)
have a microscopic expression in terms of Kubo relations
and low-frequency response functions as they can be
calculated in the homogeneous system at the local densities and
currents. Such a general view is reflected by the present result.
In the Tomonaga-Luttinger-Liquid free harmonic Hamiltonian,
where the interactions are effectively embodied in $u$ and $K$, the
Navier-Stokes equations become indeed the simple hydrodynamic relations of
Sec.~\ref{sec:hydro}. On the other hand, in the DFT and LDA approach of
Sec.~\ref{sec:approx} the treatment explicitly uses the
two mappings: from interacting to non-interacting system (DFT) and
from inhomogeneous to homogeneous (LDA).
\section{1D Bose gases coupled {\it via} dipolar interactions}
~\label{sec:TLLdip}
In this Section we specialize to the case of a 1D dipolar gas in a
harmonic trapping potential. We first recall the main results known
for the homogeneous system, and then apply Eq. (\ref{eq:s-lda-harmonic})
to determine the scattered light intensity. The system is characterized by the
strength of the interactions $C_{dd}$, resulting from either magnetic
$C_{dd}=\mu_0\mu_d^2$ or electric $C_{dd}=d^2/\epsilon_0$
dipoles, where $\mu_d$ and $d$ are the magnetic and electric dipole
moments and $\mu_0$ and $\epsilon_0$ are the vacuum
permittivities. An effective Bohr radius can be defined from
$C_{dd}$ as $r_0\equiv M C_{dd}/(2 \pi \hbar^2)$ and the Hamiltonian
in effective Rydberg units $Ry^*={\hbar^2}/({2 M r_0^2})$ is
\begin{equation}
H=(n r_0)^2 \left[ -\sum_{i} \frac{\partial^2}{\partial x_i^2}+ (n
r_0) \sum_{i<j}\frac{1}{|x_i-x_j|^3} \right] \; ,
\label{eq:HdipoleRy}
\end{equation}
where lenghts are expressed in $1/n$ units.
The physics of the model is entirely specified by the dimensionless
coupling parameter $n r_0$, so that in the high-density limit the
system becomes strongly correlated and a quasi-ordered state occurs,
where the potential energy dominates.
The ground-state energy $e(n)$
of this model was determined by means of Reptation QMC method in
Ref.~\cite{ours}. In the low
$nr_0\to 0$ limit
it reproduces the Tonks-Girardeau (TG) state
energy per particle of a
free spinless Fermi gas, whose energy per particle is
$e_{TG}(n)=\pi^2(nr_0)^2/3$ $Ry^*$.
In the large
$nr_0\to \infty$ limit of high-density dipoles, it reproduces
the Dipolar Density Wave (DDW) state where
$e_{DDW}(n)=\zeta(3) (n r_0)^3$ $Ry^*$ and
$\zeta(3)=1.20205$.
The QMC thermodynamic energy per particle in Rydberg units can be
represented as an analytical function of $n r_0$:
\begin{eqnarray}
&& e_p(nr_0)=\frac{\zeta(3) (nr_0)^4+a (nr_0)^e+ b (nr_0)^f + c
(nr_0)^{(2+g)}}{1+nr_0} \\
&& +\frac{\pi^2}{3} \frac{(n r_0)^2}{1+d
(nr_0)^g}\ \nonumber \; .
\label{eq:fit}
\end{eqnarray}
The fitting coefficients, yielding a reduced $\chi^2 \simeq 5$, are:
$a =3.1(1)$, $b = 3.2(2)$, $c = 4.3(4)$, $d= 1.7(1)$, $e= 3.503(4)$,
$f= 3.05(5)$, and $g= 0.34(4)$.
The Bragg intensity is thus easily
obtained by Eq.~(\ref{eq:s-lda-harmonic}) once the value of
$\rho^\star$ is determined.
\begin{figure}[h]
\centering \includegraphics[scale=0.7]{fig2.eps}
\caption{(Color online).
1D dipolar Bose gas confined in an harmonic trap,
with $e(n)$ as determined by
QMC simulations.
$S(q,\omega)$ (arbitrary units) {\it vs.} $\omega/(u_0 q)$
for different densities $n_0$ at the trap center.
The values of $n_0 r_0$ running from 0.01 to 1000 are indicated in the legend.}
\label{fig:scaled-plot}
\end{figure}
In Fig.\ref{fig:scaled-plot} we report the scaling behavior of
$S(q,\omega)$ {\it vs.} $\omega/(u_0 q)$ for different densities at the
trap center $n_0$. Larger $n_0 r_0$ indicate stronger coupling
interactions, crossing over from TG to DDW states.
The linear behavior in the low $\omega/q$ regime is
striking, the slope continuously increasing with decreasing $u(0)$ and
thus $n_0$. In the TG limit, the tail of
$S(q,\omega)$ is insensitive to changes of the density at the
center of the trap, and in fact the curves with $n_0r_0=0.01$ and
$0.1$ do coincide. The comparison with the TG gas ($\gamma=1$ and
$n_0r_0=0.01$) and the DDW case ($\gamma=2$ and
$n_0r_0=10^3$) is better seen in Fig.\ref{fig:crossover}, where
$S(q,\omega) \sqrt{e'(n_0)}/n_0$ is plotted as a function of
$\omega/(u_0 q)$. One can notice that a crossover takes place in the
intermediate densities regime. Viewed in the log-log scale, the plot
evidences how a measure of
the $S(q,\omega)$ tail towards small $\omega/q$, would
provide a way to determine the interaction regime. A peculiarity of
the TLL behavior is the power-law trend when $\omega/(u_0q=1)$
is approached. A detailed study of the power-law non-analyticity for a
trapped Bose gas can be found in
~\cite{golovach_2009}.
\begin{figure}[h]
\centering \includegraphics[scale=0.7]{fig3.eps}
\caption{The same as Fig.~\ref{fig:scaled-plot}, but in a log-log
scale. The
comparison with the Tonks-Girardeau limit
gas ($\gamma=1$) and the dense dipole
limit corresponding to a Dipolar Density Wave ($\gamma=2$) is shown
in evident manner.}
\label{fig:crossover}
\end{figure}
\section{Conclusions}~\label{sec:conc}
We have derived the dynamical structure factor for an
inhomogeneous Tomonaga-Luttinger liquid as it can occur in a confined
strongly interacting one-dimensional gas. In view of current
experimental progress in the field, we have provided an
easy-to-use and simple analytical expression for the light-scattering cross
section, Eq. (\ref{eq:s-lda-master}), valid within a Local Density
Approximation.
The analytical expression (\ref{eq:s-lda-master})
predicts that $S(q,\omega)$ is only a function of
$\omega/|q|$ and is the central result of this work.
In fact, this is the specific signature of
Tomonaga-Luttinger Liquid behavior in shallow trapped 1D Bose systems, along
with a power-law behavior when $\omega/(u_0 q)$ is approached, as it
can be measured by Bragg spectroscopy.
Expression (\ref{eq:s-lda-master}) is validated by the
independent derivation (\ref{eq:final-boso}) by means of a
hydrodynamic approach, which is reported in detail in
App.~\ref{sec:sums}. The connection between the two approaches is
a second result of this work, and is a consequence of the more
general Current-Density
Functional Theory~\cite{Vignale97,Chiofalo98} applied to the conditions
of the present work.
We thus remark that expression (\ref{eq:s-lda-master}) can be in principle
applied to the many 1D systems cited in the introductory
material, once the trapping potential is known together with
the ground state energy as a
function of the density, {\it e.g.} by means of perturbative, exact, or
computational methods applied to the homogeneous system. Extension of
the present method to include additional local perturbations coupling
to the density, could be used to investigate the propagation of local
density fluctuations.
Finally, we have applied our findings to the case of
one-dimensional quantum bosonic gases with dipolar interactions,
using the harmonic profile typical of
experiments in this field, accompanied by
our previous QMC
data for the energy per particle. We find an universal
scaling behavior peculiar of the Tomonaga-Luttinger
liquid~\cite{ours}, a signature that
can be eventually probed by Bragg spectroscopy in ongoing
experimental realizations of such systems~\cite{Chotia_2011}.
|
\section{Extremality in terms of complementary channels}
In what follows $\mathcal{H}$ (possibly with indices) denotes a separable
Hilbert space, $\mathfrak{T}(\mathcal{H})$ -- the Banach space of
trace-class operators and $\mathfrak{B}(\mathcal{H})=\mathfrak{T}(\mathcal{H}
)^{\ast }$-- the algebra of all bounded operators in $\mathcal{H}$. Let $A,B$
be two quantum systems with the Hilbert spaces $\mathcal{H}_{A},\mathcal{H}
_{B},$ which we call the input and the output systems. In this paper we call
by \textit{channel} a normal, unital, completely positive map $\Phi :
\mathfrak{B}(\mathcal{H}_{B})$ $\longrightarrow \mathfrak{B}(\mathcal{H}
_{A}).$ There is a unique linear trace-preserving map $\Phi _{\ast }:
\mathfrak{T}(\mathcal{H}_{A})$ $\longrightarrow \mathfrak{T}(\mathcal{H}
_{B}) $ \ such that $\Phi =\left( \Phi _{\ast }\right) ^{\ast },$ which maps
density operators (states) in $\mathcal{H}_{A}$ into density operators in $
\mathcal{H} _{B}.$ In physical terms we work in the Heisenberg picture,
while $\Phi _{\ast }$ is the channel in the Schr\"{o}dinger picture.
The set of all channels with input $A$ and output $B$ is convex. We first
give a convenient formulation of necessary and sufficient condition for a
channel to be extreme point of this set in terms of complementary channel, a
notion of big importance in quantum information theory \cite{ds,h,mr}. This
formulation is based on the general approach to extremality of completely
positive maps of a $C^*$-algebra due to Arveson \cite{ar} (see also
Appendix). We then apply this formulation to prove the main result of this
note: under certain nondegeneracy conditions, purity of the environment is
necessary and sufficient for extremality of Bosonic linear (quasi-free)
channel. It follows that a Gaussian channel between finite-mode Bosonic
systems is extreme if and only if it has minimal noise. For channels in one
Bosonic mode this was conjectured by Ivan, Sabapathy and Simon \cite{ivan}
basing on consideration of finite-dimensional criterion of Choi \cite{choi}.
Finding the proof for this statement was the initial motivation of the
present work.
Given three quantum systems $A,B,E$ with the spaces $\mathcal{H}_{A},
\mathcal{H}_{B},\mathcal{H}_{E}$ and an isometric operator $V:\mathcal{H}
_{A}\rightarrow \mathcal{H}_{B}\otimes \mathcal{H}_{E}$, the relations
\begin{eqnarray}
\Phi \lbrack X] &=&V^{\ast }(X\otimes I_{E})V,\qquad X\in \mathfrak{B}(
\mathcal{H}_{B}), \label{compl} \\
\quad \tilde{\Phi}[Y] &=&V^{\ast }(I_{B}\otimes Y)V,\qquad Y\in \mathfrak{B}
( \mathcal{H}_{E}) \label{compl1}
\end{eqnarray}
define two channels $\Phi :\mathfrak{B}\left( \mathcal{H}_{B}\right)
\rightarrow \mathfrak{B}\left( \mathcal{H}_{A}\right) ,$ $\tilde{\Phi}:
\mathfrak{B}\left( \mathcal{H}_{E}\right) \rightarrow \mathfrak{B}\left(
\mathcal{H}_{A}\right) ,$ which are called mutually \textit{complementary}.
The Stinespring dilation theorem implies that for given a channel $\Phi $
the representation (\ref{compl}) and hence a complementary channel $\tilde{
\Phi}$ given by (\ref{compl1}) always exists. The representation (\ref{compl}) is \textit{minimal} if the subspace
\begin{equation*}
\mathcal{M}=\{(X\otimes I_{E})V\psi :\psi \in \mathcal{H}_{A},X\in
\mathfrak{ \ \ \ \ B}(\mathcal{H}_{B})\}\subset
\mathcal{H}_{B}\otimes \mathcal{H}_{E}
\end{equation*}
is dense in $\mathcal{H}_{B}\otimes \mathcal{H}_{E}.$ Let
\begin{equation}
\Phi \lbrack X]=V^{\prime \ast }(X\otimes I_{E^{\prime }})V^{\prime },\qquad
X\in \mathfrak{B}(\mathcal{H}_{B}) \label{sties1}
\end{equation}
be another representation for $\Phi ,$ then there exists an isometric
operator $W$ from $\mathcal{H}_{E}$ into $\mathcal{H}_{E^{\prime }}$ such
that
\begin{equation}
(I_{B}\otimes W)V=V^{\prime }, \label{WVtildeV}
\end{equation}
so that the new complementary channel is $\tilde{\Phi}^{\prime }[Y^{\prime
}]=\tilde{\Phi}[W^{\ast }Y^{\prime }W].$ In particular, if the
representation (\ref{sties1}) is also minimal, the operator $W$ maps $
\mathcal{H}_{E}$ onto $\mathcal{H}_{E^{\prime }},$ so that the minimal
representation and the corresponding complementary channel are unique up to
the unitary equivalence. In this case the complementary channel is also
called minimal.
Let $Y$ be an operator in $\mathcal{H}_{E}$ such that $0\leq Y\leq I_{E},$
then the relation
\begin{equation*}
\Phi _{Y}[X]=V^{\ast }(X\otimes Y)V,\qquad X\in \mathfrak{B}(\mathcal{H}
_{B}),
\end{equation*}
apparently defines a completely positive map satisfying $\Phi _{Y}\leq \Phi
. $ The operator-algebraic version of the Radon-Nikodym theorem established
by Arveson \cite{ar}, Theorem 1.4.2, implies: assuming that the
representation (\ref{compl}) is minimal, this relation sets one-to-one
affine correspondence between the order intervals $[0,I_{E}]$ and $[0,\Phi
]. $
\textbf{Proposition 1.} \textit{Channel $\Phi $ is extreme if and only if it
has a complementary channel $\tilde{\Phi}^{\prime }$ such that $\mathrm{Ker}
\tilde{\Phi}^{\prime }=0,$ or, equivalently, $\overline{\mathrm{Ran}\tilde{
\Phi}_{\ast }^{\prime }}=\mathfrak{T}(\mathcal{H}_{E}).$ Any such
complementary channel is minimal.}
This follows from Theorem 1.4.6 \cite{ar} (see Appendix) but we give
here a proof for completeness. Assume that $\mathrm{Ker}\tilde{\Phi}
^{\prime }=0,$ and let $\tilde{\Phi}$ be a complementary channel in
the minimal representation (\ref{compl}) for $\Phi ,$ so that
$\tilde{\Phi} ^{\prime }[Y^{\prime }]=\tilde{\Phi}[W^{\ast
}Y^{\prime }W]$ where $W^{\ast }W=I_{E}.$ Then $\tilde{\Phi}^{\prime
}[I_{E^{\prime }}-WW^{\ast }]=\tilde{ \Phi} [W^{\ast }\left(
I_{E^{\prime }}-WW^{\ast }\right) W]=0,$ therefore $ I_{E^{\prime
}}-WW^{\ast }=0,$ so that $W$ is unitary operator onto $
\mathcal{H}_{E^{\prime }}.$ Therefore we can assume
$\tilde{\Phi}^{\prime }= \tilde{\Phi}$ and
$\mathrm{Ker}\tilde{\Phi}=0.$ Let $\Phi =\frac{1}{2}\left( \Phi
_{1}+\Phi _{2}\right) ,$ where $\Phi _{j}$ are channels. Then by
Arveson's Theorem, $\Phi _{j}=\Phi _{Y_{j}};j=1,2,$ where $0\leq
Y_{j}\leq 2I_{E}.$ Since $\Phi _{j}$ are unital, $V^{\ast
}(I_{B}\otimes Y_{j})V=V^{\ast }(I_{B}\otimes I_{E})V,$ whence
$V^{\ast }(I_{B}\otimes \left( I_{E}-Y_{j}\right) )V=0.$ But this
means that $\tilde{\Phi} [I_{E}-Y_{j}]=0,$ hence $I_{E}-Y_{j}=0$.
Therefore $\Phi _{j}=\Phi ,$ so $ \Phi $ is extreme channel.
Conversely, let $\Phi $ be extreme channel and let us show that $\mathrm{Ker}
\tilde{\Phi}=0$ for the minimal complementary channel. Let $\tilde{\Phi}
[Y]=0.$ Without loss of generality we can assume that $Y=Y^{\ast }$ and $
\left\Vert Y\right\Vert \leq 1.$ Define $\Phi _{\pm }[X]=V^{\ast }(X\otimes
\left( I_{B}\pm Y\right) )V,$ for the minimal representation (\ref{compl})
of the channel $\Phi .$ Then $\Phi _{\pm }$ are normal completely positive
maps by Arveson's Theorem and $\Phi _{\pm }[I_{B}]=I_{A\text{ }}$by the
assumption. Since $\Phi $ is extreme, $\Phi =\Phi _{\pm },$ implying $
V^{\ast }(X\otimes Y)V=0$ for all $X,Y.$ It follows that $\left\langle \psi
_{2}|V^{\ast }(X_{2}^{\ast }\otimes I_{E})(I_{B}\otimes Y)(X_{1}\otimes
I_{E})V|\psi _{1}\right\rangle \equiv 0,$ hence the bilinear form of the
operator $I_{B}\otimes Y$ vanishes on $\mathcal{M}$, hence $Y=0$ by the
minimality of the representation.
\textbf{Remark.} Choosing an orthonormal basis $\left\{ e_{j}\right\} $ in $
\mathcal{H} _{E},$ introduce the operators $V_{j}:\mathcal{H}_{A}$ to $
\mathcal{H}_{B}$, defined by $V_{j}={\langle }e_{j}|\,V$ i.e.
\begin{equation}
{\langle }\phi |\,V_{j}\psi {\rangle }={\langle }\phi \otimes
e_{j}|\,V\psi { \rangle },\quad \phi \in \mathcal{H}_{B},\psi \in
\mathcal{H}_{A}, \label{braA}
\end{equation}
and let $y_{jk}={\langle }e_{j}|\,Ye_{k}{\rangle }$ be the matrix of the
operator $Y.$ Then the above result amounts to the following: \textit{\
channel $\Phi $ is extreme if and only if it has a representation}
\begin{equation}
\Phi [X]=\sum_{k}V_{k}^{\ast }XV_{k},\qquad X\in \mathfrak{B}( \mathcal{H}
_{B}), \label{kr}
\end{equation}
\textit{where the system $\left\{ V_{j}^{\ast }V_{k}\right\} $ is strongly
independent in the sense that $\sum_{jk}y_{jk}V_{j}^{\ast }V_{k}=0$ (strong
operator convergence) for a matrix $\left[ y_{jk}\right] $ of bounded
operator implies $y_{jk}\equiv 0$.} This result contained in \cite{ts}
generalizes Choi's criterion for finite dimensional case \cite{choi}.
However for our purposes the formulation in terms of the complementary
channel turns out to be more convenient.
\section{Extremality of Linear Bosonic Channels}
In what follows we consider Bosonic system with $s$ modes described by
irreducible Weyl-Segal system
\begin{equation}
W(z)=\exp i\,(Rz), \label{1.1}
\end{equation}
in a Hilbert space $\mathcal{H}$, where
\begin{equation}
Rz=\sum_{j=1}^{s}(x_{j}q_{j}+y_{j}p_{j}),
\end{equation}
so that $R=[q_{1},p_{1},\dots ,q_{s},p_{s}]$ is the row vector of the
\textit{canonical observables } and $z=[x_{1},y_{1},\dots
,x_{s},y_{s}]^{\top }$ is the column vector of real parameters. We refer to
\cite{ho1,dvv,cgh,ssa} for relevant definitions and results.
A real vector space $Z$ equipped with a nondegenerate antisymmetric
bilinear form $\Delta $ is called symplectic space. Particularly
important case is the standard symplectic space $Z=\mathbb{R}^{2s}$
equipped with the form $ z^{\top }\Delta z^{\prime },$ where
\begin{equation}
\Delta =\left[
\begin{array}{ccccc}
0 & -1 & & & \\
1 & 0 & & & \\
& & \ddots & & \\
& & & 0 & -1 \\
& & & 1 & 0
\end{array}
\right] \label{delta}
\end{equation}
is the matrix of commutators of the canonical observables, $[Rz,Rz^{\prime
}]=-iz^{\top }\Delta z^{\prime }I.$
The noncommutative Fourier transform of a trace class operator
$\tau$ in $ \mathcal{H}$ is defined as
\begin{equation*}
\phi _{\tau}(z)=\mathrm{Tr}\tau W(z).
\end{equation*}
The complex function $\phi _{\tau}(z)$ is
bounded and continuous on $Z.$ If $\rho $ is a density operator (state), $\phi _{\rho}$ is called its
characteristic function. Then $\phi _{\rho}(0)=1$. Operator $\tau$ is positive if and only if $\phi
_{\tau}(z)$ is $\Delta - $\textit{nonnegative definite}: all the matrices
with the elements
\begin{equation}
\phi _{\tau}(z_{r}-z_{s})\,\exp \frac{i}{2}\bigl(z_{r}^{\top }\Delta
z_{s} \bigr) , \label{pd}
\end{equation}
where $z_{1},\ldots ,z_{n}$ is an arbitrary finite subset of $Z$, are
nonnegative definite.
Let $(Z_{A},\Delta _{A}),(Z_{B},\Delta _{B})$ be the symplectic spaces of
dimensionalities $2s_{A},2s_{B},$ which will describe the input and the
output of the channel (here $\Delta _{A},\Delta _{B}$ have the canonical
form (\ref{delta})) and let $W_{A}(z_{A}),W_{B}(z_{B})$ be the Weyl
operators in the Hilbert spaces $\mathcal{H}_{A},\mathcal{H}_{B}$ of the
corresponding Bosonic systems. Channel $\Phi $ transforming the Weyl
operators according to the rule
\begin{equation}
\Phi \lbrack W_{B}(z_{B})]=W(Kz_{B})f(z_{B}), \label{linbos}
\end{equation}
where $K$ is a linear map between output and input symplectic spaces and $f$
is a complex continuous function such that $f(0)=1$, is called \textit{linear Bosonic} \cite{hol} or
quasi-free \cite{dvv}. Define the real skew-symmetric $2s_{B}\times 2s_{B}-$
matrix
\begin{equation*}
\Delta _{K}=\Delta _{B}-K^{\top }\Delta _{A}K.
\end{equation*}
The map (\ref{linbos}) is completely positive if and only if $f$ is $\Delta
_{K}-$nonnegative definite: all the matrices with the elements
\begin{equation}
f(z_{r}-z_{s})\,\exp \frac{i}{2}\bigl(z_{r}^{\top }\Delta _{K}z_{s}\bigr),
\label{twcp}
\end{equation}
where $z_{1},\ldots ,z_{n}$ is an arbitrary finite subset of $Z$, are
nonnegative definite \cite{dvv},\cite{hw}. In what follows we assume that
\textit{\ the real skew-symmetric $2s_{B}\times 2s_{B}-$matrix $\Delta
_{K}=\Delta _{B}-K^{\top }\Delta _{A}K$ is nondegenerate, i.e.}
\begin{equation}
\det \Delta _{K}\neq 0. \label{nondeg}
\end{equation}
Under this condition there exist real nondegenerate $2s_{B}\times 2s_{B}-$
matrix $K_{D}$ such that
\begin{equation}
\quad K_{D}^{\top }\Delta _{D}K_{D}=\Delta _{K}, \label{iskom2}
\end{equation}
where $\Delta _{D}=\Delta _{B}.$ This is just the canonical form of the
nondegenerate skew-symmetric matrix $\Delta _{K}$. Comparing (\ref{twcp})
with (\ref{pd}) we find that there exists a state $\rho _{D}$ of the Bosonic
system in the space $\mathcal{H}_{D}$ corresponding to the standard
symplectic space $(Z_{D},\Delta _{D})\simeq (Z_{B},\Delta _{B})$ such that
\begin{equation}
f(z_{B})=\mathrm{Tr}\rho _{D}W_{D}(K_{D}z_{B})=\phi _{\rho _{D}}(K_{D}z_{B})
\label{iskom}
\end{equation}
i.e.
\begin{equation}
\Phi \lbrack W_{B}(z_{B})]=W(Kz_{B})\phi _{\rho _{D}}(K_{D}z_{B}).
\label{qfc}
\end{equation}
The relation (\ref{iskom2}) implies that
\begin{equation}
\Delta _{B}=K^{\top }\Delta _{A}K+K_{D}^{\top }\Delta _{D}K_{D}. \label{com}
\end{equation}
We will make use of the unitary dilation of the channel $\Phi $ from
\cite {cgh}. Consider the composite Bosonic system $AD=BE$ with the
Hilbert space $ \mathcal{H}_{A}\otimes \mathcal{H}_{D}\simeq
\mathcal{H}_{B}\otimes \mathcal{ H}_{E}$ corresponding to the
symplectic space $Z=Z_{A}\oplus Z_{D}=Z_{B}\oplus Z_{E},$ where
$(Z_{E},\Delta _{E})\simeq (Z_{A},\Delta _{A})$. Thus
$[R_{A}\,R_{D}]=[R_{B}\,R_{E}]$ describe different splits of the set
of canonical observables for the composite system. The channel $\Phi
$ is described by the linear input-output relation (preserving the
commutators as follows from (\ref{com})
\begin{equation}
R_{B}^{\prime }=R_{A}K+R_{D}K_{D} \label{ior}
\end{equation}
where the system $D$ is in the state $\rho _{D}$ (for simplicity of
notations we write $R_{A},\dots $ instead of $R_{A}\otimes I_{D},\dots $).
It is shown that the commutator-preserving relation (\ref{ior}) can be
complemented to the full linear canonical transformation by putting
\begin{equation}
R_{E}^{\prime }=R_{A}L+R_{D}L_{D}, \label{iocomp}
\end{equation}
where $\left( 2s_{A}\right) \times \left( 2s_{A}\right) -$ matrix
$L$ and $ \left( 2s_{B}\right) \times \left( 2s_{A}\right) -$ matrix
$L_{D}$ are such that the $2\left( s_{A}+s_{B}\right) \times 2\left(
s_{A}+s_{B}\right) -$ matrix
\begin{equation}
T=\left[
\begin{array}{cc}
K & L \\
K_{D} & L_{D}
\end{array}
\right] \label{bltr}
\end{equation}
is symplectic, i.e. satisfies
\begin{equation*}
T^{\top }\left[
\begin{array}{cc}
\Delta _{A} & 0 \\
0 & \Delta _{D}
\end{array}
\right] T=\left[
\begin{array}{cc}
\Delta _{B} & 0 \\
0 & \Delta _{E}
\end{array}
\right]
\end{equation*}
and $\Delta _{E}=\Delta _{A}.$
\textbf{Lemma 1.} \textit{Under the condition (\ref{nondeg}), $\det
L\neq 0.$ }
\textit{Proof.} The fact that $T$ is symplectic implies, in addition
to (\ref {com}),
\begin{eqnarray*}
\quad 0 &=&K^{\top }\Delta _{A}L+K_{D}^{\top }\Delta _{D}L_{D}, \\
\Delta _{E} &=&L^{\top }\Delta _{A}L+L_{D}^{\top }\Delta _{D}L_{D}.
\end{eqnarray*}
Taking into account that $\det K_{D}\neq 0,$ the first equation implies
\begin{equation*}
L_{D}=-\left( K_{D}^{\top }\Delta _{D}\right) ^{-1}K^{\top }\Delta _{A}L.
\end{equation*}
Substituting into the second equation gives $\Delta _{D}=L^{\top }ML,$ where
$M=\Delta _{A}+\Delta _{A}K\left( K_{D}\Delta _{D}\right) ^{-1}\Delta
_{D}\left( K_{D}^{\top }\Delta _{D}\right) ^{-1}K^{\top }\Delta _{A}.$
Therefore $\left( \det L\right) ^{2}\det M=1,$ whence $\det L\neq 0.\square $
Denote by the $U_{T}$ the unitary operator in $\mathcal{H}_{A}\otimes
\mathcal{H}_{D}\simeq \mathcal{H}_{B}\otimes \mathcal{H}_{E}$ implementing
the symplectic transformation $T$ so that
\begin{equation}
\lbrack R_{B}^{\prime }\,R_{E}^{\prime }]=U_{T}^{\ast
}[R_{B}\,R_{E}]U_{T}=[R_{A}\,R_{D}]T. \label{deistvo}
\end{equation}
Then we have the unitary dilation
\begin{equation}
\Phi \lbrack W_{B}(z_{B})]=\mathrm{Tr}_{D}\left( I_{A}\otimes \rho
_{D}\right) U_{T}^{\ast }\left( W_{B}(z_{B})\otimes I_{E}\right) U_{T}.
\label{udi1}
\end{equation}
The \textit{weakly complementary} channel \cite{cgh} is then
\begin{equation*}
\tilde{\Phi}^{w}[W_{E}(z_{E})]=\mathrm{Tr}_{D}\left( I_{A}\otimes \rho
_{D}\right) U_{T}^{\ast }\left( I_{B}\otimes W_{E}(z_{E})\right) U_{T}.
\end{equation*}
The equation (\ref{iocomp}) is nothing but the input-output relation for the
weakly complementary channel which thus acts as
\begin{equation}
\tilde{\Phi}^{w}[W_{E}(z_{E})]=W_{A}(Lz_{E})\phi _{\rho _{D}}(L_{D}z_{E}).
\label{Gc}
\end{equation}
In the case of pure state $\rho _{D}=|\psi _{D}\rangle \langle \psi
_{D}|$ the relation (\ref{udi1}) amounts to the Stinespring
representation (\ref {compl}) for the channel $\Phi $ with the
isometry $V=U_{T}|\psi _{D}\rangle ,$ and the relation (\ref{udi1})
amounts to (\ref{compl1}) implying that $
\tilde{\Phi}^{w}=\tilde{\Phi}.$
Apparently if the channel $\Phi $ given by (\ref{qfc}) is extreme then $\rho
_{D}$ is a pure state (otherwise the spectral decomposition of $\rho _{D}$
would provide a nontrivial convex decomposition of $\Phi $). In the converse
direction we prove
\textbf{Theorem.} \textit{Assume that} $\rho _{D}$ \textit{is a pure
state with nonvanishing characteristic function } $\phi _{\rho
_{D}}$ \textit{ which is } $ L^{2}-$\textit{differentiable to the
order} $2s_{D}$. \textit{Then the channel $\Phi $ given by
(\ref{qfc}) is extreme.}
\textbf{Remark.} We conjecture that a similar result should hold without
assumption (\ref{nondeg}) for a Bosonic linear channels on the CCR-algebra.
In \cite{dvv} purity of $\rho _{D}$, in the situation where $K$ is
symplectic transformation or symplectic projection (so that $\det \Delta
_{K}=0$), was shown to be sufficient for extremality of a quasi-free map on
the CCR-algebra.
\textit{Proof.} We shall show that the complementary channel $\tilde{\Phi}$
satisfies the condition $\overline{\mathrm{Ran}\tilde{\Phi}_{\ast }}=
\mathfrak{T}(\mathcal{H}_{E})$ of the Proposition 1.
\textbf{Lemma 2.} \textit{Let }$\mathit{\Psi _{K,f}}$\textit{\ be a Bosonic
linear channel with the same input and output space $\mathcal{H}$,}
\begin{equation*}
\Psi_{K,f} \lbrack W(z)]=W(Kz)f(z),
\end{equation*}
\textit{\ where $K$ is a nondegenerate square matrix. Then the restriction
of $\Psi _{K,f}$ onto $\mathfrak{T}(\mathcal{H})$ coincides with $\left\vert
\det K\right\vert ^{-1}\left( \Psi _{\hat{K},\hat{f}}\right) _{\ast },$
where $\hat{K}=K^{-1},\hat{f}(z)=f(-K^{-1}z).$}
\textit{Proof.} We use the inversion formula for the noncommutative Fourier
transform \cite{ho1}: if $\tau$ is trace-class operator and $\phi _{\tau}(z)=\mathrm{Tr}
\tau W(z), $ then
\begin{equation*}
\tau=\frac{1}{\left( 2\pi \right) ^{s}}\int \phi _{\tau}(z)W(-z)d^{2s}z.
\end{equation*}
It follows that
\begin{eqnarray*}
\Psi _{K,f}[\tau] &=&\frac{1}{\left( 2\pi \right) ^{s}}\int \phi
_{\tau}(z)W(-Kz)f(-z)d^{2s}z \\
&=&\frac{1}{\left( 2\pi \right) ^{s}\left\vert \det K\right\vert }\int \phi
_{\tau}(K^{-1}z)W(-z)f(-K^{-1}z)d^{2s}z \\
&=&\frac{1}{\left( 2\pi \right) ^{s}\left\vert \det K\right\vert }\int \phi
_{\tau}(\hat{K}z)W(-z)\hat{f}(z)d^{2s}z.
\end{eqnarray*}
Then
\begin{eqnarray*}
\mathrm{Tr}\Psi _{K,f}[\tau]W(z) &=&\left\vert \det K\right\vert ^{-1}\phi
_{\tau}( \hat{K}z)\hat{f}(z) \\
&=&\left\vert \det K\right\vert ^{-1}\mathrm{Tr}\tau W(\hat{K}z)\hat{f}(z) \\
&=&\left\vert \det K\right\vert ^{-1}\mathrm{Tr}\tau\Psi _{\hat{K},\hat{f}
}(W(z))
\end{eqnarray*}
and the Lemma is proved.$\square $
\textbf{Lemma 3.} \textit{Under the condition of Lemma 2,
}$\overline{ \mathit{\mathrm{Ran}\tilde{\Phi}_{\ast
}}}=\mathfrak{\ T}(\mathcal{H}_E).$
\textit{Proof.} From Lemma 2 it follows that up to a positive factor
$\tilde{ \Phi}_{\ast }$ is itself Bosonic linear channel satisfying
the condition of Lemma 2. It is sufficient to prove that arbitrary
positive trace-class operator $\tau$ in $\mathcal{H}_E$ can be
approximated in the trace norm by operators of the form $\Phi _{\ast
}[\tau_{n}],\tau_{n}\in \mathfrak{T}( \mathcal{H}_A ).$ By the
Parceval identity for the noncommutative Fourier transform, the
function $\phi _{\sqrt{\tau}}(z)$ is square integrable. Denote by
$\mathcal{C}$ the class of infinitely differentiable functions with
finite support. Let $\left\{ \phi _{n}(z)\right\} \subset
\mathcal{C}$ be a sequence converging to $\phi _{\sqrt{\tau}}(z)$ in
$L^{2}(Z).$ Consider the operators
\begin{equation*}
\sigma_{n}=\frac{1}{\left( 2\pi \right) ^{s}}\int \phi _{n}(z)W(-z)d^{2s}z.
\end{equation*}
We can assume that $\phi _{n}(-z)=\overline{\phi _{n}(-z)}$ so that
$\sigma_{n}$ is Hermitean. By the Parceval identity for the
noncommutative Fourier transform, the Hilbert-Schmidt norms
$\left\Vert \sqrt{\tau}-\sigma_{n}\right\Vert _{2}\longrightarrow
0.$ By using the inequality $\left\Vert AB\right\Vert _{1}\leq
\left\Vert A\right\Vert _{2}\left\Vert B\right\Vert _{2}$ we
conclude $ \left\Vert \tau-\sigma_{n}^{2}\right\Vert
_{1}\longrightarrow 0.$ The function $\varphi
_{n}(z)=\mathrm{Tr}\sigma_{n}^{2}W(z)$ is twisted convolution of two
functions $\phi _{n}(z)$ and hence also belongs to $ \mathcal{C}$.
Then by change of variables
\begin{eqnarray*}
\sigma_{n}^{2}&=&\frac{1}{\left( 2\pi \right) ^{s}}\int \varphi
_{n}(z)W(-z)d^{2s}z \\
&=& \frac{1}{\left( 2\pi \right) ^{s}\left\vert \det L\right\vert }\int
\varphi _{n}(L^{-1}z)W(-L^{-1}z)d^{2s}z=\tilde{\Phi}_{\ast }[\tau_{n}],
\end{eqnarray*}
where $\tau_{n}=\frac{1}{\left( 2\pi \right) ^{s}}\int \frac{\varphi
_{n}(L^{-1}z)}{f(L^{-1}z)}W(-z)d^{2s}z$ and $f(z)$ is given by
(\ref{iskom}), so it does not vanish and is $ L^{2}-$differentiable
to the order $2s$ by the assumption of Theorem. Hence the function
$\frac{\varphi _{n}(L^{-1}z)}{ f(L^{-1}z)}$ is well defined,
finitely supported and also $ L^{2}-$differentiable to the order
$2s$. It remains to show that $\tau_{n}$ is a trace class operator.
\textbf{Lemma 4.} \textit{If $\phi _{\tau }$ has finite support and
is }$ L^{2}-$\textit{differentiable of the order} $2s$
\textit{\ then $\tau \in
\mathfrak{T}(\mathcal{H}).$}
\textit{Proof.} By using the formula (see \cite{ho1}, Lemma V.4.2),
\begin{equation*}
\phi _{\tau (Rw)}(z)=\left[ -\frac{1}{2}z^{\top }\Delta w-\nabla _{w}\right]
\phi _{\tau }(z),\quad w,z\in Z,
\end{equation*}
we see that $\phi _{\tau (Rw)^{2s}}$ is square integrable. It
follows that $ \tau (Rw)^{2s}$ extends to a Hilbert-Schmidt
operator, and similarly the operator $\sigma =\tau \left(
2N_{1}+1\right) \dots \left( 2N_{s}+1\right) ,$ where
$2N_{j}+1=\left( Re_{j}\right) ^{2}+\left( Rh_{j}\right)
^{2}=q_{j}^{2}+p_{j}^{2}\ $for a symplectic basis $\left\{
e_{j},h_{j}\right\} _{j=1,\dots ,s}$ in $Z.$ Here $N_{j}$ is the
number operator of $j-$th mode which is selfadjoint with the
eigenvalues $ n_{j}=0,1,\dots ,$ and the operators $N_{1},\dots
,N_{s}\ $commute. Therefore $\sigma ^{\ast }\sigma $ is a positive
trace class operator. From this we conclude
\begin{equation*}
\mathrm{Tr}\sigma ^{\ast }\sigma =\sum\limits_{n_{1},\dots n_{s}}\left(
2n_{1}+1\right) ^{2}\dots \left( 2n_{s}+1\right) ^{2}\langle n_{1},\dots
n_{s}|\tau ^{\ast }\tau |n_{1},\dots n_{s}\rangle <\infty ,
\end{equation*}
where $\left\{ |n_{1},\dots n_{s}\rangle \right\} $ is the orthonormal basis
of common eigenvectors of operators $N_{1},\dots ,N_{s}.$ By using
Cauchy-Schwarz inequality and the inequality $\langle \psi |\,|\tau |\,|\psi
\rangle ^{2}\leq \langle \psi |\tau ^{\ast }\tau |\psi \rangle $ for a unit
vector $\psi ,$ we have
\begin{eqnarray*}
\left( \mathrm{Tr}|\tau |\right) ^{2} &=&\left( \sum\limits_{n_{1},\dots
n_{s}}\langle n_{1},\dots n_{s}|\,|\tau |\,|n_{1},\dots n_{s}\rangle \right)
^{2} \\
&\leq &\sum\limits_{n_{1},\dots n_{s}}\left( 2n_{1}+1\right)
^{2}\dots \left( 2n_{s}+1\right) ^{2}\langle n_{1},\dots n_{s}|\tau
^{\ast }\tau |n_{1},\dots n_{s}\rangle
\\&\cdot & \sum\limits_{n_{1},\dots n_{s}}\left( 2n_{1}+1\right)
^{-2}\dots \left( 2n_{s}+1\right) ^{-2}<\infty ,
\end{eqnarray*}
Thus Lemma 4 and hence Lemma 3 are proved. Applying the Proposition 1 proves
the Theorem.$\square $
\section{\protect\bigskip The case of Gaussian channels}
The density operator (state) $\rho $ is called \textit{\ Gaussian}, if its
characteristic function $\phi _{\rho }(z)=\mathrm{Tr}\rho W(z)$ has the form
\begin{equation}
\phi _{\rho }(z)=\exp \left( il^{\top }z-\frac{1}{2}z^{T}\alpha z\right) ,
\label{GaussianState}
\end{equation}
where $\alpha $ is a real symmetric $(2s)\times (2s)$-matrix, called
covariance (or correlation) matrix of $\rho $. The necessary and sufficient
condition for $\alpha $ to be a covariance matrix is the inequality
\begin{equation}
\alpha \geq \frac{i}{2}\Delta , \label{n-s condition}
\end{equation}
where both parts are considered as complex Hermitian matrices. This is
equivalent to the $\Delta -$nonnegative definiteness of $\phi _{\rho }(z).$
\textbf{Proposition 2.} \textit{Gaussian state $\rho $ is pure if and only
one of the following equivalent conditions holds:}
\begin{enumerate}
\item \textit{$\alpha $ is a minimal (in the sense of partial ordering of
real symmetric matrices) solution of the inequality (\ref{n-s condition});}
\item \textit{the symplectic eigenvalues of the matrix $\alpha $ are all
equal to their minimal possible value $\frac{1}{2};$}
\item $\mathrm{Rank}\left( \alpha -\frac{i}{2}\Delta \right) =s;$
\item $\alpha +\frac{1}{4}\Delta \alpha ^{-1}\Delta =0$;
\item \textit{$\alpha =-\frac{1}{2}\Delta J$, where $J$ is an operator of
complex structure in $(Z,\Delta )$.}
\end{enumerate}
This statement is a collection of results scattered in literature; its proof
is essentially based on Williamson's canonical form of the matrix $\alpha $,
see \cite{ho1,lin,ssa}.
\bigskip \textit{Bosonic Gaussian channel} $\Phi =\Phi _{K,l,\mu }$ is
Bosonic linear channel with Gaussian function $f(z),$ namely
\begin{equation}
\Phi _{K,l,\mu }[W_{B}(z_{B})]=W(Kz_{B})\exp \left[ il^{\top }z_{B}-\frac{1}{
2}z_{B}^{\top }\mu z_{B}\right] , \label{bosgaus}
\end{equation}
Here $K$ is real $2s_{A}\times 2s_{B}-$matrix and $\mu $ is real
symmetric $ 2s_{B}\times 2s_{B}-$matrix satisfying the inequality
\begin{equation}
\mu \geq \frac{i}{2}\left[ \Delta _{B}-K^{\top }\Delta _{A}K\right] ,
\label{nis}
\end{equation}
which is necessary and sufficient condition for complete positivity. The
channel (\ref{bosgaus}) has \textit{minimal noise} if $\mu $ is a minimal
solution of this inequality \cite{lin,cgh}.
Assuming the condition (\ref{nondeg}), let $K_{D}$ be a solution of
(\ref {iskom2}), then $\alpha _{D}=\left( K_{D}^{\top }\right)
^{-1}\mu \left( K_{D}\right) ^{-1}\ $is real symmetric $2s_{B}\times
2s_{B}-$matrix such that $\alpha _{D}\geq \frac{i}{2}\Delta _{D}.$
Further, from minimality of $ \mu $ it follows that $\alpha _{D}$ is
a minimal solution of the inequality $ \alpha _{D}\geq
\frac{i}{2}\Delta _{D},$ and as such it is the covariance matrix of
a pure centered $(l=0)$ Gaussian state $\rho _{D}=|\psi _{D}\rangle
\langle \psi _{D}|$ of the Bosonic system in the space
$\mathcal{H}_{D}$ corresponding to the standard symplectic space
$(Z_{D},\Delta _{D})\simeq (Z_{B},\Delta _{B}).$ Applying
Theorem 1 to the case of Gaussian $\rho _{D}$ we
obtain
\textbf{Corollary.} \textit{Bosonic Gaussian channel is extreme if and only
if it has the minimal noise.}
Lemma 2 in the case of Gaussian channels amounts to the statement:
\textit{Let $\Phi _{K,l,\mu }$ be the Gaussian channel with the same input
and output space, where $K$ is a nondegenerate square matrix. Then the
restriction of $\Phi _{K,l,\mu }$ onto $\mathfrak{T}( \mathcal{H})$
coincides with $\left\vert \det K\right\vert ^{-1}\left( \Phi _{\hat{K},\hat{
l},\hat{\mu}}\right) _{\ast },$ where}
\begin{equation*}
\left( \hat{K},\hat{l},\hat{\mu}\right) =\left( K^{-1},-\left( K^{-1}\right)
^{\top }l,\left( K^{-1}\right) ^{\top }\mu K^{-1}\right) .
\end{equation*}
This generalizes the duality observed for the one-mode Gaussian channels in
the canonical form in \cite{ivan}.
\section{Appendix}
\bigskip Let $\mathfrak{B}$ be a $C^{\ast }-$algebra and $\Phi $ a
completely positive map from $\mathfrak{B}$ to $\mathfrak{B}(\mathcal{H}),$
where $\mathcal{H}$ is a separable Hilbert space. Let
\begin{equation*}
\Phi \lbrack X]=V^{\ast }\pi (X)V,\quad X\in \mathfrak{B},
\end{equation*}
be a minimal Stinespring representation for $\Phi ,$ where $V$ is a
bounded operator from $\mathcal{H}$ to $\mathcal{K}$, another
separable Hilbert space, and $\pi $ is a representation of
$\mathfrak{B}$ on $\mathcal{K}$. Let $\mathfrak{E}=\pi
(\mathfrak{B})^{\prime }$ be the commutant of the algebra $\pi
(\mathfrak{B})$ in $\mathcal{K}$. Theorem 1.4.6 in \cite{ar} says
that $\Phi $ \textit{is an extreme point of the convex set of
completely positive maps }$\Psi $\textit{\ normalized so that }$\Psi
\lbrack I]= N\equiv V^{\ast }V$\textit{\ if and only if the subspace
}$\mathcal{L} =[V \mathcal{H}]\subseteq \mathcal{K}$\textit{\ is
separating for }$ \mathfrak{E}$ \textit{, i.e. for }$Y\in
\mathfrak{E}$ \textit{the relation } $P_{
\mathcal{L}}Y|_{\mathcal{L}}=0$\textit{\ implies }$Y=0. $
Consider the map
\begin{equation*}
\tilde{\Phi}[Y]=V^{\ast }YV,\quad Y\in \mathfrak{E},
\end{equation*}
which is a completely positive map satisfying $\tilde{\Phi}[I]=N,$ which may
be called \textit{complementary} to $\Phi $ (this possibility of
generalizing the notion of complementary map was noticed by Ruskai \cite{rus}
). Then it is easy to see that the above Arveson's criterion of extremality
is equivalent to $\mathrm{Ker}\tilde{\Phi}=0.$ \bigskip
\textbf{Acknowledgments.} The author is grateful to participants of the seminar
``Noncommutative Probability, Statistics and Information'' at the Steklov
Mathematical Institute for useful discussion. He acknowledges partial support of RFBR
grant and of the RAS program ``Mathematical control theory''.
\newpage
|
\section{Introduction}
A wide variety of observations point to the existence of dark matter. Several different candidate
particles have been proposed with masses varying from $10^{-6}$eV for Axions to $10^{16}$GeV
WIMPzillas, as well as much more massive composite objects. Maybe the best-motivated candidate is the
Weakly Interacting Massive Particle (WIMP) with a mass $m_{\chi}={\cal O}(\rm{GeV -TeV})$. One reason is
that the so-called hierarchy problem of the standard model suggests the existence of an additional particle
with a mass around the weak scale. Such a weak scale particle, if produced thermally, would have an
abundance today, similar to the measured dark matter density \cite{Lee:1977ua}. The relic abundance
after statistical\footnote{We distinguish here between statistical (chemical) freeze-out of
number-changing reactions, and kinetic (thermal) freeze-out.}
freeze-out depends on the particle mass, the freeze-out temperature, and the annihilation cross section. The
freeze-out temperature is related to the particle mass, $T_{\rm{fo}}\simeq m_{\chi}/25$.
Below this temperature, WIMP production ceases and the WIMPS react only kinetically
with the remaining Standard-Model particles. The kinetic freeze-out is delayed and happens at
a temperature of a few MeV \cite{Green:2005fa}.
Assuming that all dark matter is made of WIMPS, without asymmetry between WIMPs and anti-WIMPs,
and that there are no complications such as co-annihilation, one can use the WMAP 7 yr constraint on
the relative relic abundance
\cite{WMAP7yr}
\begin{equation}
\Omega_{\rm{DM}}h^2 =0.1109 \pm 0.0056
\end{equation}
to fix the annihilation cross-section as a function of WIMP mass $m_{\chi}$. Several observations of candidate
dark-matter annihilation products, like $e^{\pm}$, (anti-)protons, (anti-)deuterons, (anti-)neutrinos and
photons constrain the dark matter cross-sections. For a recent overview see for example
\cite{Catena:2009tm}. The standard calculations, as described above, assume no asymmetry between neutrinos and anti-neutrinos.
In this paper, we investigate how a neutrino asymmetry affects the WIMP abundance.
Dropping the assumption of equal numbers of neutrinos and anti-neutrinos leads to
additional contribution to the total energy density in the early universe and thus potentially alters the WIMP
relic abundance. This slight modification of the Standard Model, which might emerge from a fundamental
model of neutrino masses and interactions, is well within observational limits and might play a crucial role in
determining the relation between the relic abundance of a WIMP and its annihilation cross-section.
Very little is known about neutrino (flavour) asymmetries, due to the lack of any direct observation of cosmological neutrinos and the large number of different theories of lepto- and baryogenesis. Thus, the cosmic background of neutrinos may hide a large neutrino-anti-neutrino asymmetry, orders of magnitude larger than the baryon asymmetry \cite{Simha:2008mt}, $b\equiv\frac{n_b-n_{\bar{b}}}{s}\simeq {\cal O}(10^{-10})$.
Here $n_b$ and $n_{\bar{b}}$ denote the number densities of baryons and anti-baryons,
respectively, while $s$ is the entropy density.
All experimental bounds on the neutrino (flavour) asymmetries are inferred indirectly from measurements of
the primordial abundances of light elements, the cosmic microwave background (CMB), and large scale structure (LSS) and combinations of them.
When the universe was roughly $0.1$ seconds old and had a temperature of
a few MeV, the neutrinos were in thermal equilibrium with the photons and $e^{\pm}$ via processes like $e^+ + e^- \leftrightarrow \nu_i + \bar{\nu}_{i}$, with $i=e$,$\mu$,$\tau$. Subsequent cooling of the universe allowed the neutrinos to kinetically decouple from the plasma. As muon and tau neutrinos interact via neutral currents only, they decouple first. The electron neutrinos interact further via charged-current interactions with the ambient electrons, neutrons and protons. For this reason, an asymmetry in the electron neutrinos influences the neutron-to-proton ratio at weak-interaction-freeze-out, and thereby the $^4$He-abundance. Additionally any lepton flavour asymmetry could be detected through its additional contribution to the total energy density of the universe, often expressed as the number of extra effective
relativistic degrees-of-freedom of the plasma
\begin{equation}
\Delta N_{{\rm eff}}=\sum_{f = e,\mu,\tau}\left[\frac{30}{7}\left( \frac{\mu_{\nu_{f}}}{\pi T}\right)^2+\frac{15}{7}\left( \frac{\mu_{\nu_{f}}}{\pi T}\right)^4 \right],
\end{equation}
where $\mu_{\nu_{f}}$ are the chemical potentials of the three neutrino flavour.
To be more specific, the neutrino flavour asymmetry influences the $^4$He abundance in two ways:
Increasing $\mu_{\nu_e}$ would lead to a smaller $n/p$ fraction at the freeze out of the weak interaction
rates and hence to a smaller $^4$He abundance. But this effect can be compensated by increasing
simultaneously $|\mu_{\nu_{\mu}}|$ and/or $|\mu_{\nu_{\tau}}|$. This would raise the expansion rate and
thus lead to a higher freeze out temperature for the weak interactions and increase the $^4$He abundance.
Note, that the chemical potentials in $\Delta N_{\rm{eff}}$ enter quadratically and thus a
measurment of $\Delta N_{\rm{eff}}$ does not constrain the sign of the individual flavour asymmetries.
Thus even a vanishing net lepton asymmetry might give rise to interesting and non-vanishing lepton-flavour
asymmetries, i.e. $l=\sum_f l_f=0$. The neutrino
flavour asymmetries can be played against each other to match the observational bounds on light element
abundances \cite{Olive:1991ru}.
After the epoch of Big Bang Nucleosynthesis (BBN), the total energy density of the universe consisted of photons
and the three types of relativistic neutrinos. Altering that energy density with a large $\Delta N_{{\rm eff}}$,
e.g. large $\mu_{\nu_f}/T$, would affect the CMB spectrum by changing the redshift of matter
radiation equality and by changing the amount of anisotropic stress \cite{Hou:2011ec}.
The power spectrum of density fluctuations on small scales would be suppressed, leading to observable
effects in the large scale structure \cite{Hou:2011ec}.
The BBN bounds on the pseudo-chemical potentials
$\xi_f := \mu_{\nu_f}/T$ and on $\Delta N_{\rm{eff}}$ depend on assumptions regarding the efficiency of
neutrino flavour equilibration via neutrino oscillations.
Before the onset of neutrino flavour oscillations (i.e. at $T>T_{\rm{osc}}$), each individual lepton flavour,
\begin{equation}
l_f\equiv\frac{n_f-n_{\bar{f}}+n_{\nu_f}-n_{\nu_{\bar{f}}}}{s},
\end{equation}
is conserved.
After the onset of neutrino flavour oscillations, only the total lepton number $l\equiv\sum_{f=e,\mu,\tau}l_f$
remains conserved,
while the individual flavour asymmetries vary.
It might be that $\nu$-oscillations ensure the full equilibration of three initially different flavour asymmetries
\cite{Dolgov:2002ab, Abazajian:2002qx, Wong:2002fa, Mangano:2011ip},
such that the asymmetry in the electron neutrinos measured today, would be the same as in the muon and
tau type.
In \cite{Dolgov:2002ab,Wong:2002fa}, it is shown that even large primordial asymmetries with
$\mu_{\nu_e}/T=0$, $\mu_{\nu_\mu}/T=-0.1$, and $\mu_{\nu_\tau}/T=0.1$ equilibrate at roughly
$T\simeq 4$ MeV.
Thus, one can assume that, at least for $\mu_{\nu_f}/T = {\cal O}(0.1)$,
neutrino flavour equilibration happens at $T \sim 10$ MeV, well before the onset of BBN \cite{Mangano:2011ip}.
As a consequence LSS, CMB, and BBN are blind to the earlier differences in the individual lepton flavour.
For this case ($l = \sum_f l_f = 3 l_e$), the observation of
primordial abundances and WMAP data constrain all three pseudo-chemical potentials:
$\xi_f \leq 0.023\pm 0.041$ \cite{Simha:2008mt},
assuming the lepton asymmetry to be constant between nucleosynthesis and photon decoupling.
The recently reported He$^4$ abundance by the ACT collaboration \cite{Dunkley:2010ge} and the WMAP7 data release lead to a significantly larger range
$-0.14 \leq \xi_f \leq 0.12$ \cite{Krauss:2010xg}.
Considering only partial equilibration of three initially different flavour asymmetries via oscillations
(i.e.~$\mu_{\nu e} \neq \mu_{\nu_\mu} \neq \mu_{\nu_\tau}$), weakens the bounds on $\xi_f$.
Assuming $\mu_{\nu_e} \ll \mu_{\nu_f}$ and an effective number of neutrinos
$N_{\rm eff}=3.3^{+0.7}_{-0.6}$ while neglecting flavour equilibration,
leads to the bound $|\xi_f|\leq 2.34$ (here $f = \mu,\tau$)
and a total lepton asymmetry of $|l| \leq 5$ \cite{Simha:2008mt}. We conclude, large lepton asymmetries
and large lepton flavour asymmetries before the onset of neutrino oscillations are compatible with current observations.
The consequences of lepton asymmetries on the cosmic QCD
transition at $T_{\rm qcd} \sim 200$ MeV are discussed in \cite{Schwarz:2009ii}.
There it is shown that large neutrino asymmetries influence
significantly the dynamics of the QCD transition. As the electric charge, baryon and lepton
flavour are preserved at the same time, a non-negligible lepton asymmetry can induce a non-vanishing
baryon-chemical potential, which in turn influences the pseudo-critical temperature or even turns a crossover transition into a first order transition.
In this work we investigate an even earlier event: statistical
freeze-out of the abundance of weakly interacting dark matter.
We show that large lepton (flavour-) asymmetries ($|l| > 0.01$) give rise to sizeable effects on the relic abundance of WIMP dark matter.
In the following chapter we we will give an (incomplete) overview of arguments how large lepton (flavour) asymmetries can be generated. In the third chapter we show then, how large lepton (flavour) asymmetries influences the thermodynamic description of the early universe. We give analytic estimations and numerical results and show how these effect the relative relic abundance of WIMP dark matter in the fourth chapter before we conclude our work in chapter 5.
Throughout the article we set $c=\hbar=k_B=1$.
\section{A short review of models with large lepton (flavour) asymmetry}
Several ideas of baryo- and leptogenesis have been proposed
that lead to a universe with large lepton flavour asymmetries but small baryon asymmetry.
For example, in grand unified theories asymmetries $|{l}|\gg b$ can be realized based on gauge invariant initial particle asymmetries \cite{Harvey:1990qw}.
Another mechanism is provided by supersymmetric theories, c.f.~\cite{McDonald:1999in},
which have the intriguing feature of flat directions. Made up of squark or slepton fields,
these directions can carry baryon and/or lepton number. During cosmological inflation these
squark and sleptons are free to fluctuate and to form scalar condensates, carrying baryon
and/or lepton number, and to release these charges by decaying to Standard-Model particles.
This is called the Affleck-Dine mechanism \cite{Affleck:1984fy,Dolgov:1990zm},
and in principle can give rise to large asymmetries in baryons or leptons.
In the early universe, large asymmetries have to be produced after an inflationary phase,
otherwise they diluted away by the inflationary expansion and washed out
by the huge increase in entropy density during post-inflationary reheating.
Another possibility to dilute a large lepton (flavour) asymmetry are sphaleron transitions at high temperatures. These transitions violate $b$ and $l$, but preserve $b-l$. Sphalerons are the reason for suggesting ${\cal O}(b)\simeq {\cal O}(l)$. Note that this does not necessarily mean
${\cal O}(b)\simeq {\cal O}(l_f)$. For instance, two flavours could be orders of magnitude larger, if
their sum almost cancels, $l_i \simeq - l_j$ for lepton flavours $i\neq j$.
However, sphaleron transitions might never be in equilibrium
and the argument in favor of $l = {\cal O}(b)$ would disappear.
It was shown that, if the total asymmetry $\sum_f l_f$ is larger than a critical value
$l_{\rm c}\simeq 10^{-2}$ then electroweak symmetry is never restored and sphalerons are suppressed for
all times \cite{Linde:1976kh,McDonald:1999nv}. A large lepton (flavour) asymmetry would survive
until today. It is remarkable that recent bounds on the electron-neutrino asymmetry from WMAP-7yr
alone and combined with the ACT data are in the vicinity of this limit.
In \cite{Casas:1997gx} the authors combine the Affleck-Dine mechanism with the suppressed sphalerons. They consider a model with a sneutrino condensate to generate large $l$-asymmetries, $l>l_{\rm c}$, but no baryon asymmetry. In their specific model they take the minimal supersymmetric Standard-Model with three right-handed neutrino singlet superfields. For a mass of the lightest neutrino of ${\cal O}(10^{-4} {\rm eV})$ and a preferred mass of the lightest
right-handed neutrino of ${\cal O}({\rm TeV})$,
their model leads to asymmetries $l\simeq 10^{-2}$ to 1.
In \cite{Liu:1993am,McDonald:1999nv} it is shown how the electroweak symmetry non-restoration due to a
large $l$-asymmetry works with a small baryon asymmetry.
How to regulate the baryon asymmetry in Affleck-Dine mechanisms is discussed in \cite{Campbell:1998yi}.
An interesting model is also discussed in \cite{MarchRussell:1999ig}, where $l_e=-l_{\mu}$ and $l_{\tau}=0$.
\section{Thermodynamics in the early universe}
It is an excellent approximation to assume that entropy is conserved in the early Universe, even through several phase transitions \cite{BdVS}.
Even if the QCD transition were of first order, the amount of entropy produced would be tiny \cite{SSW,BdVS}.
It is convenient to work with specific densities of conserved quantities.
The specific baryon asymmetry $b$ is related to the baryon number density $n_B$, via $b = n_B/s$,
where $s$ denotes the entropy density.
Constraints from BBN and CMB are often formulated in terms of $\eta_B = n_B/n_\gamma = b s/n_\gamma$.
Note that $\eta_B(t_{\rm bbn}) \neq \eta_B(t_{\rm cmb})$, while $b$ is constant.
Before neutrino decoupling, the radiation fluid of the early Universe is a tightly coupled plasma
in thermal and statistical equilibrium with respect to strong, electromagnetic and weak forces.
It is characterized by a single temperature $T$ and chemical potentials $\mu_i$ for each particle
species (with mass $m_i$) of the Standard-Model
($i=$ quarks/hadrons, leptons and gauge bosons and we count particles and antiparticles separately).
The chemical potentials of anti-particles $\mu_{\bar{i}}=-\mu_i$, as long
as particles are relativistic ($T > m_i/3$). (The chemical potentials of the non-relativistic particles (e.g. baryons) are
irrelevant, only because they are apparently too small for these species to contribute appreciably to the energy density, pressure and entropy of the Universe
during the epochs in question.)
In this work we consider temperatures between the electroweak and the QCD transition,
$T_{\rm ew}>T>T_{\rm qcd}$, where $T_{\rm ew}\sim 200$ GeV and $T_{\rm qcd}\sim 200$ MeV.
For WIMP masses between about $5$ GeV and $4$ TeV, it is within this temperature range
that WIMP annihilation freezes-out and the WIMP number density is determined.
The equilibrium distribution function $f_i$ of a particle $i$ and its anti-particle $\bar{i}$ is given for
bosons ($-$) and fermions ($+$) by the Bose-Einstein and Fermi-Dirac distributions,
\begin{eqnarray}
f_i (E) = \frac{1}{{\rm exp}\frac{E-\mu_i}{T} \mp 1}, \quad
f_{\bar{i}}(E) = \frac{1}{{\rm exp}\frac{E+\mu_i}{T} \mp 1},
\end{eqnarray}
where $E$ denotes the energy of the particles.
We are interested in the difference of the number densities of particles and their anti-particles,
i.e.~their net number density,
\begin{eqnarray}
\label{nint} n_i = \frac{g_i}{2\pi^2} \int^{\infty}_{m_i}{E(E^2-m_i^2)^{1/2}\left[f_i(E) - f_{\bar{i}}(E) \right]{\rm d}E},
\end{eqnarray}
where $g_i (= g_{\bar i})$ denotes the number of helicity degrees of freedom of a particle
species (e.g.~$g_e = 2$). For ultra-relativistic fermions the net number density simply becomes
\begin{equation}
n_i = \frac{g_i}{6} \mu_i T^2 \left[ 1 + \left(\frac{\mu_i}{\pi T}\right)^2 \right].
\end{equation}
All particles and anti-particles contribute to the energy density of the Universe, which is given by
\begin{eqnarray}\label{intg}
\epsilon = \sum_{j = i,\bar{i}} \frac{g_j}{2\pi^2}\int^{\infty}_{m_j}{E^2(E^2-m_j^2)^{1/2}f_j(E){\rm d}E},
\end{eqnarray}
where the index $j$ can stand for any particle or anti-particle species.
In equilibrium, electromagnetic and weak interactions provide a set of relations between the chemical potentials. It follows that all gauge bosons have vanishing chemical potentials, and for relativistic fermions
\begin{eqnarray}
\mu_i = - \mu_{\bar i},\\
\mu_u + \mu_f = \mu_d + \mu_{\nu_f},\\
\mu_u = \mu_c = \mu_t, \\
\mu_d = \mu_s = \mu_b.
\end{eqnarray}
The indices $u$, $d$, $c$, $s$, $t$ and $b$ represent the up, down, charm, strange, top and bottom
quarks. Lepton flavour ($e$, $\mu$, or $\tau$) is denoted by $f$.
This leaves us with five independent chemical potentials (i.e.~three for the neutrino flavours and two
for the up and down quarks) plus the temperature. These six variables are uniquely determined by
five conservation laws and the Friedman equations. The five conserved quantities in the energy range
of interest (between $T_{\rm ew}$ and $T_{\rm qcd}$) are electric charge ($q$),
baryon number and three lepton flavour numbers.
\begin{eqnarray}
\label{charge}s\, q &=& - \sum_{i=e,\mu,\tau} n_i + \frac 23 \sum_{i= u,c,t} n_i - \frac 13 \sum_{i = d,s,b} n_i \\
\label{baryon}s\, b &=& \frac 13 \sum_{i= u,d,c,s,t,b} n_i \\
\label{lflavour}s\, l_f &=& n_f + n_{\nu_f}, \quad f = e, \mu, \tau
\end{eqnarray}
Lepton flavour is conserved, since the timescale for $\nu$-oscillations is much larger then the
Hubble time for $T>$ few MeV. Additionally we assume that these global conservation laws hold
locally, i.e.~there are no electric currents and no baryon or lepton (flavour) diffusion.
We assume a charge neutral universe ($q=0$) and fix the baryon asymmetry to
$b = 9 \times 10^{-11}$ \cite{Simha:2008mt}. While these are well established assumptions, measurements of the lepton flavour asymmetries in the early Universe are not available.
We thus keep them as free parameters. In \cite{Schwarz:2009ii} it is shown in
detail how to extract all chemical potentials $\mu_i(T;b,\{l_f\})$, including details on the numerical
method used to solve the five conservation equations as a function of $T$.
\subsection{Effective relativistic degrees of freedom}
Let us now take a closer look at the contribution of lepton flavour asymmetries on the effective
relativistic degrees of freedom contributing to the total energy density
\begin{equation}
g_{\ast}(T,\{\mu_i\}){\equiv}\frac{30}{\pi^2T^4} \epsilon(T,\{\mu_i\}).
\end{equation}
Together with the solution $\mu_i = \mu_i(T;b,l_e, l_\mu, l_\tau)$ we find
\begin{equation}
g_{\ast}=g_{\ast}(T;b,l_e,l_{\mu},l_{\tau}).
\end{equation}
In the ultra-relativistic case ($m_i=0$) and for vanishing baryon and lepton asymmetry (i.e.~vanishing chemical potentials) we recover
\begin{equation}
g_{\ast}(T,\{0\})=\sum_{i = {\rm bosons}} g_i + \frac{7}{8}\sum_{i={\rm fermions}} g_i.
\end{equation}
Large lepton flavour asymmetries generically lead to large chemical potentials of all fermion species.
For ultra-relativistic fermions, (\ref{intg}) can be solved exactly. Assuming $m_i = 0$ and
$\mu_i=-\mu_{\bar{i}}$ leads for the energy density of a fermion species to
\begin{equation}
\epsilon_{i + \bar{i}} \stackrel{\mu_i=-\mu_{\bar{i}}}{=} \frac{\pi^2}{30} T^4 (2 g_i)
\left[\frac{7}{8} + \frac{15}{4} \left(\frac{\mu_i}{\pi T}\right)^2
+ \frac{15}{8} \left(\frac{\mu_i}{\pi T}\right)^4 \right].
\end{equation}
This leads to an increase of $g_\ast$, due to non-vanishing chemical potentials,
\begin{eqnarray}
\label{gast}
g_{\ast} (T,\{\mu_i\})
&=g_{\ast}(T,{0}) + \Delta g_{\ast}(T,\{\mu_i\}),
\end{eqnarray}
with
\begin{equation}\label{Deltagast}
\Delta g_{\ast} (T,\{\mu_i\}) =
\sum_i g_i \left[\frac{15}{4} \left(\frac{\mu_i}{\pi T}\right)^2 +
\frac{15}{8} \left( \frac{\mu_i}{\pi T}\right)^4 \right].
\end{equation}
Any nonzero $\Delta g_{\ast}$ would therefore increase the total energy density and thus the Hubble expansion rate.
This increased expansion rate would alter the relic abundances of light elements, and so must be checked against observations.
The increased abundance of relativistic species is normally expressed as an increase of the effective number of neutrinos $N_{\nu_{\rm{eff}}}$.
>From the LEP measurement of the decay width of the $\rm{Z}^0$ boson, we expect three active neutrino species with masses well below the electroweak scale, thus
\begin{equation}
N_{\nu_{\rm{eff}}}= 3 + \frac{30}{7}\sum_{f}\left[\left(\frac{\mu_{\nu_f}}{\pi T}\right)^2 +
\frac{1}{2}\left(\frac{\mu_{\nu_f}}{\pi T} \right)^4\right].
\end{equation}
We also have to take a closer look at the $\mu$-dependence of $s/T^3$. In kinetic theory, including quantum statistics, the entropy density in terms of the distribution function $f$ is \cite{Bernstein}
\begin{equation}\label{sint}
s = -\int{\left[f {\rm ln}f \mp(1 \pm f){\rm ln}(1 \pm f)\right]\frac{{\rm d}^3p}{(2\pi)^3}},
\end{equation}
where upper and lower signs refer to boson and fermion statistics, respectively.
For ultra-relativistic particles, vanishing chemical potentials and all particle species at temperature
$T$, we recover the well known result
\begin{eqnarray}\label{entropyT}
s(T) \stackrel{m,\mu=0}{=} && \frac{2\pi^2}{45}T^3g_{\ast}.
\end{eqnarray}
Taking chemical potentials into account leads to extra contributions
\begin{eqnarray}\label{entropymu}
s(T,\mu_i)&\stackrel{m=0}{=}&
\frac{2\pi^2}{45} T^3 \left[g_\ast + \frac{15}{8}\sum_{i={\rm fermions}}g_i\left(\frac{\mu_i}{\pi T}\right)^2\right] \\
&=&\frac{2\pi^2}{45}T^3\left( g_{\ast}+\Delta g_{s_{\ast}}\right).
\end{eqnarray}
Note that $\Delta g_{s_{\ast}} \neq \Delta g_{\ast}$ and, in contrast to the energy density, terms quartic in the chemical potentials do not show up in the entropy density. When $T \sim m_i$, we have to
calculate the entropy density numerically.
\subsection{Analytic estimates and numerical results}
For an analytic estimate of the effect of lepton flavour asymmetries, we neglect all masses
of quarks and leptons and assume that all lepton flavour asymmetries are small enough to
justify $\mu_i/(\pi T) \ll 1$. Further we assume $m_W/3 > T > m_b/3$, thus $g_\ast = 345/42$.
{}From the conservation of charge, baryon number and lepton flavour we find
\begin{eqnarray}
0 &=& \frac 1 3 T^2 (4 \mu_u - 3 \mu_d - \mu_e - \mu_\mu - \mu_\tau) + {\cal O}(\mu_i^3), \\
\frac{23 \pi^2}{6} T^3 b &=& \frac{1}{3} T^2 (2 \mu_u + 3 \mu_d) + {\cal O}(b \mu_i^2, \mu_i^3), \\
\frac{23 \pi^2}{6} T^3 l_f &=& \frac 1 6 T^2 (2 \mu_f + \mu_{\nu_f}) + {\cal O}(l_f \mu_i^2,\mu_i^3).
\end{eqnarray}
Solving this set of equations results in
\begin{eqnarray}
\label{mudApr}
\frac{\mu_d}{\pi T} &=& \pi \left[\frac{5}{2} b - \frac{2}{3} l \right], \\
\frac{\mu_u}{\pi T} &=& \pi \left[2 b + l \right], \\
\frac{\mu_f}{\pi T} &=& \pi \left[\frac{1}{6} b - \frac{5}{9} l + \frac{23}{3} l_f \right], \\
\label{munuApr}
\frac{\mu_{\nu_f}}{\pi T} &=& \pi \left[- \frac{1}{3} b + \frac{10}{9} l + \frac{23}{3} l_f \right].
\end{eqnarray}
Lepton (flavour) densities large compared to the baryon density affect not only the number densities of leptonic species, but also those of quarks. This can lead to increase of the
effective degrees of freedom in the energy density and the entropy density. Note that all relativistic particle species equipped with baryon or lepton number contribute to this effect, as long as they are in statistical equilibrium.
\subsubsection{Flavour symmetric lepton asymmetry}
Let us first assume that all lepton flavour numbers are the same, $l_e = l_\mu = l_\tau = l/3$. We also
assume $b \ll |l| \ll 1$. Thus we can put $b=0$. This results in
\begin{equation}
\frac{\mu_d}{\pi T} = - \frac{2\pi}{3} l, \quad
\frac{\mu_u}{\pi T} = \pi l, \quad
\frac{\mu_f}{\pi T} = 2\pi l, \quad
\frac{\mu_{\nu_f}}{\pi T} = \frac{11\pi}{3}l,
\end{equation}
and allows us to estimate the change in the effective degrees of freedom,
\begin{equation}\label{gast2}
\Delta g_{\ast}(T,b=0,l_f = l/3) = \frac{1265}{2} \pi^2 l^2 \approx 6.2 \times 10^3 l^2.
\end{equation}
Thus $\Delta g_{\ast}/g_{\ast} \approx 760 l^2$, which we assumed to be small for the purpose of the analytic approximation, i.e.~$l < 10^{-2}$.
For lepton asymmetries $l > 10^{-2}$ we rely on a numerical solution of the five conservation equations and include all particles from the Standard-Model of particle physics with their measured physical
masses (the unknown masses of the Higgs and the neutrinos are irrelevant in the regime of interest).
We solved the equations (\ref{charge}) to (\ref{lflavour}) using (\ref{nint}) and (\ref{sint}) with the method described in detail in \cite{Schwarz:2009ii}.
The numerical results for the effective helicity degrees of are shown in Fig.~\ref{fig:doflsym}. We found that an asymmetry $l_f=0.01$ leads to a small deviation from the standard case with $b=l=l_f=0$. If we apply the experimentally given upper bound for the electron neutrino asymmetry to all flavour, we found for $l_f=0.1$ approximately additional 50 degrees of freedom in the early universe between $1<T<50$ GeV.
\begin{figure}[htb]
\centering
\includegraphics[angle=270,width=0.80\textwidth]{doflf.eps}
\caption{The numerical solution for the flavour symmetric case $l=3l_f$. The effective degrees of freedom of all particles in statistical equilibrium versus the temperature in GeV on logarithmic scale. The black line corresponds to standard case, where lepton asymmetries are neglected. The blue line shows the influence of $l_f=0.01$, the green $l_f=0.05$ and the red $l_f=0.1$.}
\label{fig:doflsym}
\end{figure}
\subsubsection{Flavour asymmetric lepton asymmetry}
Let us now have a closer look at scenarios in which at least one of the three flavour lepton numbers satisfies $|l_f| \gg b$, but we restrict to $|l_f| \leq 1$ for all flavour. For simplicity we can put $b=0$. The first interesting situation is that one flavour asymmetry dominates, say $l_\tau \neq 0$ and the other flavour asymmetries vanish. In that case we would find that the quark chemical potentials are affected:
\begin{eqnarray}
\frac{\mu_d}{\pi T} &=& - \frac{2\pi}{3} l_\tau, \\
\frac{\mu_u}{\pi T} &=& \pi l_\tau, \\
\frac{\mu_{e,\mu}}{T} &=& - \frac{5 \pi}{9} l_\tau, \\
\frac{\mu_\tau}{T} &=& \frac{64 \pi}{9} l_\tau, \\
\frac{\mu_{\nu_{e,\mu}}}{T} &=& \frac{10 \pi}{9} l_\tau, \\
\frac{\mu_{\nu_\tau}}{T} &=& \frac{79 \pi}{9} l_\tau.
\end{eqnarray}
The numerical results for this situation are presented in Fig.~\ref{fig:dofl=ltau}. We see again a tiny deviation from the standard case for $l_{\tau}=0.01$. For $l_{\tau}=0.1$ there would be around 10 more degrees of freedom. The effect is smaller compared to symmetric case since the total $l$ is smaller.
\begin{figure}
\centering
\includegraphics[angle=270,width=0.80\textwidth]{dofltau.eps}
\caption{The numerical solution for the flavour asymmetry $l=l_{\tau}$. The effective degrees of freedom of all particles in statistical equilibrium versus the temperature in GeV on logarithmic scale. The black line corresponds to standard case, where lepton flavour asymmetries are neglected. The green line shows the influence of $l_{\tau}=0.01$, the blue stands for $l_{\tau}=0.05$ and the red for $l_{\tau}=0.1$.}
\label{fig:dofl=ltau}
\end{figure}
Another interesting scenario is a vanishing total lepton asymmetry $l=0$, but $l_{\mu}=-l_{\tau} \neq 0$. In this case quark chemical potentials would not be affected:
\begin{equation}
\mu_d = \mu_u = \mu_e = \mu_{\nu_e} = 0, \quad
\frac{\mu_f}{\pi T} = \frac{\mu_{\nu_f}}{\pi T} = \frac{23 \pi }{3} l_f, \quad
f = \mu,\tau.
\end{equation}
We find $\Delta g_\ast = 5 (23 \pi)^2 l_\tau^2 \approx 2.6 \times 10^4 l_\tau^2$. As we assumed
for the analytic approximation that the modification is small, its regime of validity is limited to
$|l_\tau| < 10^{-2}$.
The numerical results for vanishing $l$, but non-vanishing lepton flavour asymmetry
are presented in Fig.~\ref{fig:doflmu=ltau2}. In the calculations for the degrees of freedom the sign of a possible asymmetry does not play any role, since they enter squared. For $l_{\mu}=-l_{\tau}=0.1$ additional 20 degrees of freedom appear. What makes this scenario the most interesting, is the possibility for even larger asymmetries. If we assume $l_{\mu}=-l_{\tau}=1$ we find more then 600 additional degrees of freedom.
\begin{figure}[htbp]
\centering
\includegraphics[angle=270,width=0.80\textwidth]{doflmu=-ltau2.eps}
\caption{The numerical solution for the flavour asymmetry $l_{\mu}=-l_{\tau}$ and $l=l_e=0$. The effective degrees of freedom of all particles in statistical equilibrium versus the temperature in GeV on logarithmic scale. The black line corresponds to standard case, where lepton flavour asymmetries are neglected. The green line shows the influence of $l_{\tau}=0.01$, the blue stands for $l_{\tau}=0.05$ and the red for $l_{\tau}=0.1$.}
\label{fig:doflmu=ltau2}
\end{figure}
\section{Effect on decoupling of WIMP dark matter}
The WIMP is a very well-motivated candidate to solve the dark matter puzzle \cite{Jungman:1995df}.
In the following we assume the WIMP to be the single component dark matter particle, without asymmetry between WIMPs and anti-WIMP.
In the hot early universe these particles with masses typically between 10 and 1000 GeV are in thermal and statistical equilibrium with the radiation content. Their statistical freeze-out, when WIMPs decouple statistically from the radiation plasma, happens at $T_{\rm fo} \simeq m/25$, leading to a mass dependent interval of 0.4 GeV $<T_{\rm fo}<$40 GeV (see e.g.~\cite{Green:2005fa}). In the following we want to investigate the effect of changing the Standard Model boundary condition by allowing large neutrino asymmetries on the WIMP freeze-out.
To calculate the relic density, one assumes annihilations of the WIMPs $X\bar X\rightarrow \cdots$ with a typical weak interaction cross section $\sigma \propto G_{\rm F}^2$. The corresponding
Boltzmann equation then leads to an equation for their net particle density
\begin{equation}
\dot{n} + 3Hn = -\left\langle \sigma |v| \right\rangle (n^2-n_{\rm eq}^2),
\end{equation}
with the Hubble parameter $H$, the total annihilation cross section $\sigma$ and the relative velocity
of the annihilating particles $v$. The index ``eq'' indicates the assumption of
thermal and statistical equilibrium.
Since the WIMP is non relativistic at the time of decoupling, the annihilation cross section can be
approximated for most cases as $\langle \sigma v \rangle=
a + b/x + {\cal O}(x^{-2})$,
where the numbers $a$ and $b$ describe s- and
p-wave annihilation and $x \equiv m/T$, for more details see e.g.~\cite{Jungman:1995df}.
It is common to introduce the specific WIMP abundance $Y = n/s$ and to assume the conservation
of entropy:
It is convenient to analyze the freeze-out process in a comoving volume, thus we introduce the specific WIMP abundance $Y = n/s$ and $Y_{\rm eq} = n_{\rm eq}/s$ and the ratio $x \equiv m/T$:
\begin{equation}
\frac{{\rm d} Y}{{\rm d}x}=-\frac{\langle \sigma |v| \rangle s}{3c_s^2 H x}(Y^2 - Y_{\rm eq}^2).
\end{equation}
In radiation domination $3 c_s^2 = 1$ is
a good approximation.
As freeze-out happens when the rate of annihilations drops below the Hubble rate, we find
$Y_{\infty} \approx Y_{\rm fo} \approx [H/(\langle \sigma v\rangle s)]_{\rm fo}$.
Thus an approximate solution for the WIMP relic abundance today can be derived as
\begin{equation}
Y_{\infty}\equiv Y(x\rightarrow \infty) \approx \left(\frac{\sqrt{g_\ast}}{g_{S\ast}}\right)_{\rm fo}
\frac{1}{m_{\rm Pl} m (a/x_{\rm{fo}} + b/x^2_{\rm{fo}})}.
\end{equation}
for more details see e.g.~\cite{Lee:1977ua,Drees:2007kk}.
It is convenient to express the relative WIMP mass density
$\Omega_{\rm wimp} = \rho_{\rm wimp}/\rho_{\rm c}$ with the critical density $\rho_{\rm c}$.
The present relic abundance is then $\rho_{\rm wimp} =
m n_{0}= m s_0Y_{\infty}$, where the subscript zero denotes the value today:
\begin{eqnarray}\label{Omegachi}
\Omega_{\rm wimp}h^2 &=&
2.7 \times 10^{10}\frac{m}{100 \rm{GeV}} Y_{\infty} \\
&\simeq& \frac{8.5 \times 10^{-11}}{\rm{GeV}^2} \left(\frac{\sqrt{g_\ast}}{g_{S\ast}}\right)_{\rm fo}
\frac{x_{\rm{fo}}}{a + b/x_{\rm{fo}}}.
\end{eqnarray}
The abundance of a WIMP particle is inverse proportionally to its annihilation cross section, a more strongly interacting particle stays longer in equilibrium. The dependence on helicity degrees of freedom is apparent in the denominator, but there is also an implicit dependence in
$x_{\rm fo} \approx c - \ln(g_{\ast})$, where the constant c includes a logarithmic dependence
on the wimp mass, its cross section and its helicity degrees of freedom \cite{Green:2005fa}.
As long as $\Delta g_\ast \ll g_\ast$, which is the case for $l_f \ll 1$, we find
\begin{equation}
\label{Omega}\frac{\Delta \Omega_{\rm wimp}}{\Omega_{\rm wimp}} =
\frac 12 \left( 1 - (1 + \ell) \frac 1{x_{\rm fo}} \right)
\frac{\Delta g_\ast}{g_\ast} - \frac{\Delta g_{S\ast}}{g_{S\ast}}
\end{equation}
where $\ell = 0,1$ when the s- or p-wave annihilation channel dominates. This can be further simplified using the expressions obtained for the ultrarelativistic degrees of freedom. In this case the contributions from $\mu^2$ terms in the first two terms cancel out,
\begin{equation}
\frac{\Delta \Omega_{\rm wimp}}{\Omega_{\rm wimp}} \sim
- (1 + \ell) \frac 1{x_{\rm fo}} \frac{15}{8} \sum_i g_i \left(
\frac{\mu_i}{\pi T}\right)^2.
\end{equation}
However, this will not be the case close to mass thresholds. For the flavour symmetric case, with
$x_{\rm fo} \approx 25$ and assuming p-wave annihilation, we find $\Delta \Omega_{\rm wimp}/
\Omega_{\rm wimp} \simeq - 30 l^2$, i.e.~a reduction of the WIMP relic density by 30\% (7\%) for
$l = 0.1 (0.05)$. However, for $l > 10^{-2}$ higher order contributions of the statistical potentials
are important and the interesting regime must be analyzed numerically.
For the numerical analysis we assume the governing effect in helicity degrees of freedom. The difference of the relative relic density introduced by lepton (flavour) asymmetries at a given freeze out temperature can be approximated
\begin{equation}
\frac{\Omega_{\rm wimp}(l,l_f\neq0)}{\Omega_{\rm wimp}(l,l_f=0)} \sim \left[\sqrt{\frac{g_{\ast}(l,l_f\neq0)}{g_{\ast}(l,l_f=0)}}~ \frac{g_{S\ast}(l,l_f=0)}{g_{S\ast}(l,l_f\neq0)}\right]_{x_{\rm{fo}}}.
\end{equation}
A numerical analysis including all mass thresholds and not restricted to small chemical potentials
is shown in figure \ref{fig:Deltamulf} for different lepton asymmetries $l_f$.
We plot the ratio $\Omega_{\rm{wimp}}(l,l_f)/\Omega_{\rm{wimp}}(l=l_f=0)$ as a function of
WIMP freeze-out temperature.
In the flavour symmetric case (figure~\ref{fig:Deltamulf}) we find
that the analytic estimate for lepton asymmetries overestimates the effect.
We observe an effect of order 1 $\%$ for $l_f=0.01$ and of almost $20\%$ for $l_f=0.1$.
\begin{figure}[htbp]
\centering
\includegraphics[angle=270,width=0.80\textwidth]{DeltaWIMPlf-1.eps}
\caption{Comparison between the relative WIMP dark matter abundance without lepton flavour asymmetries $\Omega_{DM}(l=l_f=0)$ and several large asymmetries $l_f=l_e=l_{\mu}=l_{\tau}$, plotted versus the freeze out temperature $T_{\rm{fo}}$ on a logarithmic scale. We observe for flavour asymmetries $l_f=0.05$ an effect of approximately 7 percent. For $l_f=0.1$ this effect increases to 20 percent.}
\label{fig:Deltamulf}
\end{figure}
\section{Conclusion}
In this work we extend the study of cosmic lepton and lepton flavour asymmetries to the early Universe
before the cosmic QCD transition, a subject studied previously \cite{Schwarz:2009ii}.
The standard model of cosmology does not know any lepton or baryon number violating processes after
the electroweak phase transition. A lepton flavour asymmetry produced before or during the transition
would remain constant until the onset of neutrino flavour oscillations at $T\simeq {\cal O}(10)$ MeV. Neutrino oscillations ensure that all lepton flavour asymmetries
agree at the time of BBN, i.e.~$l_e=l_{\mu}=l_{\tau}$,
and thus bounds from BBN and CMB observations apply to the total lepton asymmetry $l$ only.
These bounds are rather weak and allow $|l| \gg b$.
Before the onset of neutrino oscillations, scenarios with different individual flavour asymmetries are even
less constrained, e.g.~$l_e=b$ and $- l_{\mu} = l_{\tau} \gg b$.
In this work we took a closer look on the the freeze-out of WIMP dark matter. This event is of
particular interest, since observations of these relics would open a window to the pre-BBN era.
We have shown, that lepton asymmetries might have a sizable effect on the relic abundance of the
WIMP dark matter. The relative relic abundance depends on the annihilation cross
section, the WIMP mass and lepton flavour asymmetries.
A reduction of the WIMP abundance due to large lepton (flavour) asymmetries happened due to
an increase of the effective relativistic degrees of freedom of the radiation plasma. Large lepton
asymmetries lead generically to large chemical potentials of all fermion species, which in turn contribute
to the energy and entropy densities of the universe. We presented
analytical estimates and numerical studies of this effect. Figures
\ref{fig:doflsym} to \ref{fig:doflmu=ltau2} show how the effective relativistic degrees of freedom
$g_{\ast}$ increase for different distributions of individual lepton flavour asymmetries.
We demonstrated how a large $l_f$ influences the relic WIMP abundance in figure \ref{fig:Deltamulf}.
If the asymmetries are equal in in all flavour, $l_e=l_{\mu}=l_{\tau}$ and $|l_f|=0.1$ we found a
huge effect on the relative relic abundance of approximately 20 per cent. Even if the total
lepton asymmetry is of the order of the baryon asymmetry, but individual flavour asymmetries
are large before the onset of neutrino oscillations, a reduction of the WIMP relic density of up to
$10\%$ is possible. The effect presented in this work adds to the astrophysical and particle physics
uncertainties to be accounted for in direct and indirect dark matter searches. Large lepton
(flavour) asymmetries could be compensated by decreasing annihilation cross sections, in order
to keep the relic abundance of a WIMP candidate fixed.
\ack
We thank Tanmay Vachaspati and Yi-Zen Chu for discussions. M.S. acknowledges hospitality of Case
Western Reverse University, where part of this work was done. This work was supported by the Friedrich-
Ebert-Foundation (M.S) and Deutsche Forschungsgemeinschaft grant IRTG 881 (M.S. and D.J.S.) and by a
grant from the US-DOE to the particle astrophysics theory group at CWRU.
\section*{References}
|
\section{Introduction}
Next-Closure is one of the best known algorithms in Formal Concept Analysis~\cite{ganter1999formal}
to compute the concepts of a formal context. In its general form it is able to efficiently
enumerate the closed sets of a given closure operator on a finite set. This generality might be a
drawback concerning efficiency compared to other algorithms like
Close-by-One~\cite{conf/cla/VychodilKO10,conf/iccs/Andrews09}, but widens its field of applications.
However, there are still applications where Next-Closure might be useful, but is not applicable,
because an closure operator on a finite set is not explicitly available. One such example might be
the computation of concepts of a fuzzy formal context~\cite{DBLP:conf/icfca/BelohlavekV06}. Even
worse, if a closure operator is given, but not on a finite set, Next-Closure is not directly
applicable as well. In those cases most often an ad hoc variation of Next-Closure can be
constructed. The aim of this paper is to provide a generalization of Next-Closure which covers
those cases, and even goes beyond them.
As it turns out, Next-Closure is not about enumerating closed sets of a closure operator, even not
on an abstract ordered set. The algorithm is merely about enumerating elements of a certain
semilattice, given as an operation together with a generating set. This observation is somewhat
surprising, but, as we shall see, quite natural.
This paper is organized as follows. First of all we shall revisit the original version of
Next-Closure, together with the basic definitions. Then we present our generalized version working
on semilattices, together with a complete proof of its correctness. Then we show how this
generalized form is indeed a generalization of the original Next-Closure. Additionally, we discuss
a new algorithm for enumerating the intents of a given formal context. Finally, we give some
outlook on further questions which might be interesting within this line of research.
\section{The Next-Closure Algorithm}
Before we are going to discuss our generalized form of Next-Closure, let us revisit the original
version as it is given in~\cite{ganter1999formal,fca:Ganter:1984}. To make our discussion a bit
more consistent, we shall make one minor modifications to the presentation given here, which will be
explicitly mentioned.
Let $M$ be a finite set and let $c:\subsets{P}\to\subsets{P}$ be a function such that
\begin{enumerate}[a) ]
\item $c$ is idempotent, i.e.\ $c(c(A)) = c(A)$ for all $A\subseteq M$,
\item $c$ is monotone, i.e.\ if $A\subseteq B$, then $c(A)\subseteq c(B)$ for all $A,B\subseteq M$, and
\item $c$ is extensive, i.e.\ $A\subseteq c(A)$ for all $A\subseteq M$.
\end{enumerate}
A set $A\subseteq M$ is called \emph{closed (with respect to $c$)} if $A = c(A)$, and the
\emph{image of $c$} is defined as
\[ c[\subsets{M}] := \set{c(A)\mid A\subseteq M}. \]
Without loss of generality, let $M = \set{1,\ldots,n}$ for some $n\in\NN$. For two sets $A,B\in
c[\subsets{M}]$ with $A\neq B$ and $i\in M$ we say that $A$ is \emph{lectically smaller} than $B$
\emph{at position $i$} if and only if
\[ i = \min(A\mathrel{\Delta}B) \quad\text{and}\quad i\in B, \]
where $A\mathrel{\Delta}B = (A\setminus B)\cup(B\setminus A)$ is the symmetric difference of $A$ and
$B$. We shall write $A \prec_i B$ if $A$ is lectically smaller than $B$ at position $i$. Finally,
we say that $A$ is \emph{lectically smaller} than $B$, for $A,B\in c[\subsets{M}]$, if $A = B$ or $A
\prec_i B$ for some $i\in M$ and we shall write $A \preceq B$ in this case.
It has to be noted that, in contrast to our definition, the lectic order is normally
defined for all sets $A,B\subseteq M$ in the very same spirit as given above. However, as
we shall see, this is not necessary, wherefore we have restricted our definition to closed
sets only.
Now let us define for $A\in c[\subsets{M}]$ and $i\in M$
\[ A\oplus i := c(\set{j\in A\mid j < i} \cup \set{i}). \]
Then we have the following result.
\begin{Theorem}[Next-Closure]
Let $A\in c[\subsets{M}]$. Then the next closed set $A^+\in c[\subsets{M}]$ after $A$ with
respect to the lectic order $\preceq$, if it exists, is given by
%
\[ A^+ = A\oplus i \]
%
with $i\in M$ being maximal with $A <_i A\oplus i$.
\end{Theorem}
This is the original version of Next-Closure, as it is given in~\cite{ganter1999formal}.
Now let us have a closer look on the definition of $\oplus$. The set $A\oplus i$ can be seen as the
smallest closed set containing both $\set{j \in A\mid j < i}$ and $\set{i}$, or equivalently, both
$c(\set{j \in A\mid j < i})$ and $c(\set{i})$. This means that we can rewrite $A\oplus i$ as
\[ A\oplus i = c(\set{j \in A\mid j < i}) \vee c(\set{i}), \]
where $X\vee Y$ is the smallest closed sets containing both $X, Y\in c[\subsets{M}]$, the
\emph{supremum of $X$ and $Y$}, which is simply given by $X \vee Y = c(X\cup Y)$. This observation
allows us to consider Next-Closure on abstract algebraic structures with a binary operation $\vee$
with some certain properties. To do so we need a more general notion of $c(\set{i})$, since we do
not necessarily deal with subsets, and a more general notion of $\set{j\in A\mid j<i}$, which
likewise might not be expressible in a more general setting. Finally, we need to find a starting
point for our enumeration, which is $c(\emptyset)$ in the original description of Next-Closure, but
may vary in other cases. Luckily, all this is possible and quite natural, as we shall see in the
next section.
\section{Generalizing Next-Closure for Semilattices}
The aim of this section is to present a generalization of the Next-Closure algorithm that works on
semilattices. For this recall that a semilattice $\underline{L} = (L,\vee)$ is an algebraic
structure with a binary operation $\vee$ which is associative, commutative and idempotent. It is
well known that with
\[ x \le_{\underline{L}} y \,:\!\iff x\vee y = y, \]
with $x,y\in L$. An order relation on $L$ is defined such that for every two elements
$x,y$ the element $x\vee y$ is the least upper bound of both $x$ and $y$ with respect to
$\le_{\underline{L}}$.
For the remainder of this section let $\underline{L} = (L,\vee)$ be an arbitrary but fixed
semilattice. Furthermore, let $(x_i\mid i\in I)$ be an enumeration of a finite generating set
$\set{x_i\mid i\in I}\subseteq L$ of $\underline{L}$. Finally, let $\le_I$ be a total order on $I$.
\newcommand{\le_{\underline{L}}}{\le_{\underline{L}}}
\begin{Definition}
Let $a,b\in L$ and let $i\in I$. Set
%
\[
\Delta_{a,b} := \set{j\in I
\mid
(x_j\le_{\underline{L}} a \text{ and } x_j\not\le_{\underline{L}} b)
\text{ or }
(x_j \not\le_{\underline{L}} a \text{ and } x_j\le_{\underline{L}} b)}.
\]
%
We then define
\[
a <_i b \,:\!\iff i = \min\Delta_{a,b} \text{ and } x_i\le_{\underline{L}} b.
\]
%
Furthermore we write $a < b$ if $a <_i b$ for some $i\in I$ and write $a\le b$ if $a = b$ or $a <
b$.
\end{Definition}
One can see the similarity of this definition to the one of the lectic order. Here, the set
$\Delta_{a,b}$ generalizes $a\mathrel{\Delta}b$ and $x_i\le_{\underline{L}} b$ somehow represents the fact that
$i\in b$, or equivalently $\set{i}\subseteq b$, in the special case of $L = \subsets{M}$ and $i\in
M$.
Note that if $a <_i b$ and $k\in I$ with $k<_I i$, then
\[ x_k\le_{\underline{L}} a \iff x_k\le_{\underline{L}} b. \]
This observation is quite useful and will be used in some of the proofs later on.
The first thing we want to consider now are two easy results stating that $\le$ is a total order
relation on $L$ extending $\le_{\underline{L}}$.
\begin{Lemma}\label{lem:lectic is order}
The relation $<$ is irreflexive and transitive. Furthermore, for every two elements $a,b\in L$
with $a\neq b$, it is either $a < b$ or $b < a$.
\begin{Proof}
If $a = b$, then the set $\Delta_{a,b}$ defined above is empty, therefore we cannot have $a <_i
a$ for some $i\in I$. This shows the irreflexivity of $<$. Let us now consider the
transitivity of $<$. For this let $a,b,c\in L$, $i,j\in I$ and suppose that $a<_ib$ and $b<_j
c$. We have to show that $a < c$. Let us consider the following cases.
\textit{Case $i<_I j$.} We have $x_i\not\le_{\underline{L}} a$ and $x_i\le_{\underline{L}} b$ because of $a<_i b$. Due to $i
<_I j$ it follows that $x_i\le_{\underline{L}} c$. Suppose that there exists $k\in I$, $k <_I i$ with $x_k
\le_{\underline{L}} a$ and $x_k\not\le_{\underline{L}} c$. Then if $x_k\not\le_{\underline{L}} b$ we would have $x_k\not\le_{\underline{L}} a$ because $k
<_I i$, a contradiction. But if $x_k\le_{\underline{L}} b$, then $x_k\le_{\underline{L}} c$ because of $k <_I i <_I j$, again
a contradiction. Thus we have shown that $a < c$.
\textit{Case $j <_I i$.} We have $x_j\not\le_{\underline{L}} b$, $x_j\le_{\underline{L}} c$ because of $b <_j c$. Due to $j
<_I i$ it follows that $x_j \not\le_{\underline{L}} a$. Now if there were a $k\in I$, $k <_I j$ with $x_k\le_{\underline{L}}
a$ and $x_k\not\le_{\underline{L}} c$, then $x_k\le_{\underline{L}} b$ would imply $x_k\le_{\underline{L}} c$ and $x_k\not\le_{\underline{L}} b$ would imply
$x_k\not\le_{\underline{L}} a$, analogously to the first case, a contradiction. Hence such a $k$ cannot exist
and $a < c$.
\textit{Case $i = j$.} This cannot occur since otherwise $x_i\le_{\underline{L}} b$, because of $a<_i b$, and
$x_i\not\le_{\underline{L}} b$, because of $b<_i c$, a contradiction.
Overall we have shown that $a < c$ in any case and therefore $<$ is a transitive relation.
Finally let $a,b\in L$ with $a\neq b$. Then because $\set{x_i\mid i\in I}$ is a generating set,
the set $\Delta_{a,b}$ is not empty, since otherwise $a = b$, and with $i := \min\Delta_{a,b}$
we either have $a <_i b$ if $x_i\le_{\underline{L}} b$ and $b <_i a$ otherwise.
\end{Proof}
\end{Lemma}
\begin{Lemma}\label{lem:lectic extends original order}
Let $a,b\in L$ with $a\le_{\underline{L}} b$. Then $a\le b$. In particular, if $a \le_{\underline{L}} c$ and
$b\le_{\underline{L}} c$ for $a,b,c\in L$, then $a\vee b \le c$.
\begin{Proof}
We show $x_i\le_{\underline{L}} a\implies x_i\le_{\underline{L}} b$ for all $i\in I$. This shows $b\not<a$, hence $a\le b$ by
Lemma~\ref{lem:lectic is order}. Now if $x_i\le_{\underline{L}} a$, then because of $a\le_{\underline{L}} b$ we see that
$x_i\le_{\underline{L}} b$ and the claim is proven.
\end{Proof}
\end{Lemma}
The next step towards a general notion of Next-Closure is to provide a generalization of $\oplus$.
\begin{Definition}
Let $a\in L$ and $i\in I$. Then define
%
\[ a \oplus i := \bigvee_{\substack{j <_I i\\x_j\le_{\underline{L}} a}} x_j \vee x_i. \]
\end{Definition}
With all these definitions at hand we are now ready to prove the promised generalization.
For this, we generalize the proof of Next-Closure as it is given in~\cite[page
67]{ganter1999formal}.
\begin{Lemma}\label{lem:helper}
Let $a,b\in L$ and $i,j\in I$. Then the following statements hold:
\begin{enumerate}[i) ]
\item\label{lem:helper:1} $a <_i b, a <_j c, i <_I j \implies c <_i b$.
\item\label{lem:helper:2} $a < a \oplus i$ if $x_i\not\le_{\underline{L}} a$.
\item\label{lem:helper:3} $a <_i b \implies a\oplus i \le b$.
\item\label{lem:helper:4} $a <_i b \implies a <_i a\oplus i$.
\end{enumerate}
\begin{Proof}
\begin{enumerate}[i) ]
\item It is $x_i\not\le_{\underline{L}} a$ and due to $i<_I j$ we get $x_i\not\le_{\underline{L}} c$ as well. Furthermore,
$x_i\le_{\underline{L}} b$ because of $a <_i b$. Now if there would exist a $k\in I$ with $k <_I i$ such
that $x_k\le_{\underline{L}} c, x_k\not\le_{\underline{L}} b$, then $x_k\le_{\underline{L}} a$ because of $k <_I i$ and $x_k\not\le_{\underline{L}} a$
because of $k <_I i <_I j$, a contradiction. With the same argumentation a contradiction
follows from the assumption that there exists a $k\in I$, $k<_I i$ with $x_k\not\le_{\underline{L}} c,
x_k\le_{\underline{L}} b$. In sum we have shown $c <_i b$, as required.
\item We have $x_i\not\le_{\underline{L}} a$ and $x_i\le_{\underline{L}} a\oplus i$. Furthermore, for $k\in I$, $k <_I i$ and
$x_k\le_{\underline{L}} a$ we have $x_k\le_{\underline{L}} a\oplus i$ by definition. This shows $a < a\oplus i$.
\item Let $a <_k b$ for some $k\in I$. Then $\bigvee_{j <_I k,x_j\le_{\underline{L}} a}x_j \le_{\underline{L}} b$ and $x_k\le_{\underline{L}}
b$, hence with Lemma~\ref{lem:lectic extends original order} we get $a\oplus k\le b$.
\item Let $a <_i b$. Then $x_i\not\le_{\underline{L}} a$ and with~(\ref{lem:helper:2}) we get $a < a\oplus
i$. By~(\ref{lem:helper:3}), $a\oplus i \le b$. If for $k\in I$, $k <_I i$ it holds that
$x_k\le_{\underline{L}} a\oplus i$ and $x_k\not\le_{\underline{L}} a$, then we also have $x_k\le_{\underline{L}} a\oplus i \le_{\underline{L}} b$, i.e.\
$x_k\le_{\underline{L}} b$, contradicting the minimality of $i$.
\end{enumerate}
\end{Proof}
\end{Lemma}
\begin{Theorem}[Next-Closure for Semilattices]\label{thm:semiNC}
Let $a\in L$. Then the next element $a^+\in L$ with respect to $<$, if it exists, is given by
%
\[ a^+ = a\oplus i \]
%
with $i\in I$ being maximal with $a <_i a\oplus i$.
\begin{Proof}
Let $a^+$ be the next element after $a$ with respect to $<$. Then $a <_i a^+$ for some
$i\in I$ and by Lemma~\ref{lem:helper}.\ref{lem:helper:4} we get $a <_i a\oplus i$ and
with Lemma~\ref{lem:helper}.\ref{lem:helper:3} we see $a\oplus i \le a^+$, hence
$a\oplus i = a^+$. The maximality of $i$ follows from
Lemma~\ref{lem:helper}.\ref{lem:helper:1}.
\end{Proof}
\end{Theorem}
To find the correct element $i\in I$ such that $a^+ = a\oplus i$ can be optimized with
Lemma~\ref{lem:helper}.\ref{lem:helper:2}. Because of this result, only elements $i\in I$ with
$x_i\not\le_{\underline{L}} a$ have to be considered, a technique which is also known for the original form of
Next-Closure.
However, to make the above theorem practical for enumerating the elements of a certain semilattice,
one has to start with some element, preferably the smallest element in $L$ with respect to $\le$.
This element must also be minimal in $L$ with respect to $\le_{\underline{L}}$, by Lemma~\ref{lem:lectic extends
original order}. Since $\set{x_i\mid i\in I}$ is a generating set of $\underline{L}$, and $a \le
a \vee b$ for all $a,b\in L$, the minimal elements of $L$ with respect to $\le_{\underline{L}}$ must be among the
elements $x_i$, $i\in I$. So to find the first element of $\underline{L}$ with respect to $\le$,
find all minimal elements in $\set{x_i\mid i\in I}$ and choose the smallest element with respect to
$\le$ from them. But because of $x_i \le x_j$ if and only if $j <_I i$, one just has to take the
largest index $j$ of all minimal elements among the $x_i$ to find the smallest element in $L$ with
respect to $\le$.
As a final remark for this section note that the set $\set{x_i\mid i\in I}$ must always include the
$\vee$-irreducible elements of $\underline{L}$. These are all those elements $a\in L$ that cannot
be represented as a join of other elements, or, equivalently,
\[
\set{b \in L\mid b <_{\underline{L}} a} = \emptyset
\quad\text{or}\quad
\bigvee_{b <_{\underline{L}} a} b <_{\underline{L}} a.
\]
It is also easy to see that the $\vee$-irreducible elements of $\underline{L}$ are also sufficient,
i.e.\ they are a generating set of $\underline{L}$.
\section{Computing the Intents of a Formal Context}
We have seen an algorithm that is able to enumerate the elements of a semilattice from a given
generating set. We have also claimed that this is a generalization of Next-Closure, which we want
to discuss in this section. Furthermore, we want to give another example of an application of this
algorithm, namely the computation of the intents of a given formal context.
Firstly, let us reconstruct the original Next-Closure algorithm from Theorem~\ref{thm:semiNC} and
the corresponding definitions. For this let $M$ be a finite set and let $c$ be a closure operator
on $M = \set{0,\ldots,n-1}$, say. We then apply Theorem~\ref{thm:semiNC} to the semilattice
$\underline{P} = (c[\subsets{M}], \vee)$. We immediately see that ${\le_{\underline{P}}} =
{\subseteq}$ and that $<_i$ is the usual lectic order on $\underline{P}$. Then the set
\[ \set{c(\set{i}) \mid i\in M}\cup\set{c(\emptyset)} \]
is a finite generating set of $\underline{P}$ and we can define $x_i := c(\set{i})$ and $x_n :=
c(\emptyset)$, i.e.\ $I = \set{0,\ldots,n}$. For a closed set $A\subseteq M$ and $i\in I$ then
follows
\begin{align*}
A\oplus i
&= \bigvee_{\substack{j < i\\x_j\subseteq A}}x_i \vee x_i\\
&= c(\bigcup_{\substack{j < i\\x_j\subseteq A}} x_j) \vee x_i\\
&= c(\bigcup\set{c(\set{j})\mid j< i, j\in A}) \vee c(\set{i})\\
&= c(\set{j\mid j < i, j\in A}) \vee c(\set{i})\\
&= c(\set{j\mid j < i, j\in A} \cup \set{i})
\end{align*}
which is the original definition of $\oplus$ for Next-Closure. Furthermore, it is $A\oplus n = A$
since $c(\emptyset)\subseteq A$ for each closed set $A$. We therefore do not need to consider $x_n$
when looking for the next closed set, and indeed, the only reason why $x_n = c(\emptyset)$ has been
included is that it is the smallest closed set in $\underline{P}$. All in all, we see that
Next-Closure is a special case of Theorem~\ref{thm:semiNC}.
However, for a closure operator $c$ on a finite set $M$ it seems more natural to consider the
semilattice $\underline{P} = (c[\subsets{M}],\cap)$, because the intersection of two closed sets of
$c$ again yields a closed set of $c$. One sees that ${\le_{\underline{P}}} = {\supseteq}$. As a
generating set we take the set of $\cap$-irreducible elements $\set{X_i\mid i\in G}$ for some index
set $G$. Let $A,B\in c[\subsets{M}]$ and let $<_G$ be a linear ordering on $G$. Then $A < B$ if
and only if there exists $i\in G$ such that
\[
i = \min\set{j\in G\mid (X_i\supseteq A, X_i \not\supseteq B) \text{ or } (X_i \not\supseteq A,
X_i\supseteq B)}
\quad\text{and}\quad
X_i \supseteq B
\]
and $\oplus$ is just given by
\[ A\oplus i = \bigcap_{\substack{j <_G i\\X_j\supseteq A}}X_j\cap X_i. \]
Now note that $\oplus$ does not need the closure operator $c$ anymore. This means that if the
computation of $c$ is very costly and the $\cap$-irreducible elements (or a superset thereof) is
known, this approach might be much more efficient. In general, however, it is not known how to
efficiently determine the $\cap$-irreducible closed sets of $c$. But if $c$ is given as the
$\cdot''$ operator of a formal context, these irreducible elements can be determined
quickly~\cite{ganter1999formal}.
Let $G$ and $M$ be two finite sets and let $J\subseteq G\times M$. We then call the triple
$\mathbb{K} := (G,M,J)$ a \emph{formal context}, $G$ the \emph{objects} of the formal context and
$M$ the \emph{attributes} of the formal context. For $g\in G$ and $m\in M$ we write $g\mathrel{J}m$
for $(g,m)\in J$ and say that object $g$ \emph{has} attribute $m$.
Let $A\subseteq G$ and $B\subseteq M$. We then define the \emph{derivations} of $A$ and $B$ to be
\begin{align*}
A' &:= \set{m\in M\mid \forall g\in A: g\mathrel{J}m}\\
B' &:= \set{g\in G\mid \forall m\in B: g\mathrel{J}m}.
\end{align*}
Then the $\cdot''$ operator is just the twofold derivation of a given set of attributes. It turns
out that this is indeed a closure operator, and that every closure operator can be represented as a
$\cdot''$ operator of a suitable formal context~\cite{ganter1999formal,Borchmann11}. The closed
sets of $\cdot''$, i.e.\ all sets $B\subseteq M$ with $B = B''$, are called the \emph{intents} of
$\mathbb{K}$ and shall be denoted by $\Int(\mathbb{K})$. It is clear from the previous remarks that
$(\Int(\mathbb{K}),\cap)$ is a semilattice.
The advantage of representing a closure operator is that the $\cap$-irreducible elements of
$(\Int(\mathbb{K}),\cap)$ can be directly read off from the format context. As discussed
in~\cite{ganter1999formal}, the set
\[ \set{\set{g}'\mid g\in G} \]
contains the irreducible elements we are looking for, except $M$. Furthermore, it is possible to
omit certain objects $g$ from $\mathbb{K}$ without changing $\Int(\mathbb{K})$. Every object $g\in
G$ can be omitted from $\mathbb{K}$ for which the set $\set{g}'$ is either equal to $M$ or can be
represented as a proper intersection of other sets $\set{g_1}',\ldots,\set{g_n}'$ for some elements
$g_1,\ldots,g_n\in G$. It is also clear that if there exist two distinct objects $g_1$ and $g_2$
with $\set{g_1}' = \set{g_2}'$, that we can remove one of them without changing $\Int(\mathbb{K})$.
A formal context for which no such objects exist is called \emph{object clarified} and \emph{object
reduced}. If $\con{K} = (G,M,J)$ is an object clarified and object reduced formal context, then
the set $\set{\set{g}'\mid g\in G}$ is exactly the set of $\cap$-irreducible intents of $\con{K}$,
except for the set $M$.
The above described algorithm now takes the following form when applied to $(\Int(\con{K}),\cap)$.
As index set we choose the set $G$ of object of the given formal context, ordered by $<_G$. For
every object $g\in G$ we set $x_g := \set{g}'$. Then the set $\set{x_g\mid g\in G}$ is a generating
set of the semilattice $(\Int(\con{K})\setminus\set{M},\cap)$, which we want to enumerate (since we
get the set $M$ for free). For $A$ begin an intent of $\con{K}$ and $g\in G$ we have
\begin{align*}
A\oplus g
&:= \bigcap_{\substack{h\in G\\h<_G g}}\set{h}' \cap \set{g}'
\end{align*}
and as the first intent we take $M$. For an intent $A\subseteq M$ of $\con{K}$ we then have to find
the maximal object (with respect to $<_G$) $g\in G$ such that $A <_g A\oplus g$. This is equivalent
to $g$ being maximal with $\set{g}'\not\supseteq A$ and $\forall h\in G,h<_G g: \set{h}'\supseteq A
\iff \set{h}'\supseteq A\oplus g$. However, the direction ``$\Longrightarrow$'' is clear, hence we
only have to ensure
\[ \forall h\in G, h<_Gg: \set{h}'\not\supseteq A \implies \set{h}'\not\supseteq A\oplus g. \]
All these considerations yield the algorithm shown in Listing~\ref{alg:next closure with cap}. Of
course, the derivations of the form $\set{g}'$ should not be computed every time they are needed but
rather stored somewhere for reuse.
\begin{lstlisting}[float,caption={Compute Next Intent of a Formal Context},label={alg:next closure
with cap},language=PseudoCode,numbers=none]
define next-intent($\con{K} = (G,M,J)$,$A$)
for $g\in G$, descending
if $\set{g}'\not\supseteq A$ then
let ($B$ := $A\oplus g$)
if $\forall h\in G, h<_G g, \set{h}'\not\supseteq A: \set{h}'\not\supseteq B$ then
return $B$
end if
end let
end if
end for
return nil
end
\end{lstlisting}
In contrast to the original version of Next-Closure used to compute the intents of a given formal
context, our version only traverses the formal context $\con{K}$ once, when computing $\set{g}'$ for
every $g\in G$. This might be an advantage if accessing the context is computationally expensive,
but also be a disadvantage if there are much more objects than attributes (which occurs quite often
in practice). However, in this case one might better compute the extents of the formal context,
i.e.\ all sets $A\subseteq G$ with $A = A''$, which also form a closure system. Then the algorithm
in Listing~\ref{alg:next closure with cap} would run through the set of attributes to compute the
extent.
Finally, let us consider the time complexity of the new algorithm to compute the intent after $A$
for a given formal context $\con{K} = (G,M,J)$. If we assume the operations $\cap$ and $\supseteq$
to be constant (i.e.\ independent of the size of both $G$ and $M$, a very optimistic assumption),
then the algorithm from Listing~\ref{alg:next closure with cap} roughly needs $|G|\times(|G|+|G|)$
steps to compute the next intent, i.e.\ has worst time complexity $\mathcal{O}(|G|^2)$. On the
other hand, if one assumes a naive implementation of $\cap$ and $\supseteq$ taking time
$\mathcal{O}(|M|)$, one has time complexity $|G|\times(|M|+|G|\times|M|+|G|\times|M|)$, hence worst
case complexity $\mathcal{O}(|G|^2\times|M|)$, which is the time complexity of Next-Closure when
used to compute the extents of $\con{K}$.
\section{Conclusion}
We have seen a natural generalization of the Next-Closure algorithm to enumerate elements of a
semilattice from a generating set. We have proven the algorithm to be correct and applied it to the
standard task of computing the intents of a given formal context, yielding a new algorithm to
accomplish this. However, there are still some interesting ideas one might want to look at.
Firstly, a variation of the original Next-Closure algorithm is able to compute the stem base of a
formal context, a very compact representation of its implicational knowledge. It would be
interesting to know whether the generalization given in this paper gives more insight into the
computation, and therefore into the nature of the stem base. This is, however, quite a vague idea.
Secondly, the new algorithm discussed above to compute the intents of a formal context enumerates
them in a certain order, which might or might not be a lectic one. Understanding this order
relation might be fruitful, especially with respect to the complexity results obtained recently,
which consider enumerating pseudo-intents of a formal context in lectic
order~\cite{DBLP:conf/icfca/Distel10}.
Finally, and more technically, it might be interesting to look for other applications of our general
algorithm. As already mentioned in the introduction, the enumeration of fuzzy concepts of a fuzzy
formal context might be worth investigating (and it might be interesting comparing it
to~\cite{DBLP:journals/tfs/BelohlavekBOV10}), but there might be other applications.
\bibliographystyle{plain}
|
\section{Introduction}
From experiment it is known that the masses of the nine light
pseudo-scalar
mesons show an interesting pattern. Taking the quark model point of
view, the three lightest mesons, the pions, contain only the two
lightest quark flavours, the \emph{up}- and \emph{down}-quarks. The
pion triplet has a mass of $M_\pi\approx140\ \mathrm{MeV}$. For the other six,
the \emph{strange} quark contributes also, and hence they are
heavier. In contrast to what one might expect five of them,
the four kaons and the $\eta$ meson, have
roughly equal mass around $500$ to $600\ \mathrm{MeV}$, while
the last one, the $\eta'$ meson, is much
heavier, with mass of about $1\ \mathrm{GeV}$. On the QCD level, the
reason for this pattern is thought to be the breaking of the $U_A(1)$
symmetry by quantum effects. The $\eta'$ meson is, even in a
world with three massless quarks, not a Goldstone boson.
In this proceeding contribution, we discuss the determination of $\eta$
and $\eta'$ meson masses using twisted mass lattice QCD (tmLQCD) with
$N_f=2+1+1$ dynamical quark flavours. This will not only allow a
study of the dependence of the $\eta,\eta'$ masses on the light quark mass
value, but also an investigation of the charm quark contribution to
both of these states. Moreover, the $\eta_c$ meson mass can be studied
in principle. For recent lattice studies in $N_f=2+1$ flavour QCD
see~\cite{Christ:2010dd,Kaneko:2009za}.
\begin{table}[t!]
\centering
\begin{tabular*}{.8\textwidth}{@{\extracolsep{\fill}}lccccc}
\hline\hline
ensemble & $\beta$ & $a\mu_\ell$ & $a\mu_\sigma$ & $a\mu_\delta$ &
$L/a$ \\
\hline\hline
A40.24 & $1.90$ & $0.0040$ & $0.150$ & $0.190$ & $24$ \\
A60.24 & $1.90$ & $0.0060$ & $0.150$ & $0.190$ & $24$ \\
A80.24 & $1.90$ & $0.0080$ & $0.150$ & $0.190$ & $24$ \\
\hline
A80.24s & $1.90$ & $0.0080$ & $0.150$ & $0.197$ & $24$ \\
\hline
B25.32 & $1.95$ & $0.0025$ & $0.135$ & $0.170$ & $32$ \\
B35.32 & $1.95$ & $0.0035$ & $0.135$ & $0.170$ & $32$ \\
B85.24 & $1.95$ & $0.0085$ & $0.135$ & $0.170$ & $24$ \\
\hline\hline
\end{tabular*}
\caption{The ensembles used in this investigation. The notation of
ref.~\cite{Baron:2010bv} is used for labeling the ensembles.}
\label{tab:setup}
\end{table}
\section{Lattice Action}
We use gauge configurations as produced by the European Twisted Mass
Collaboration (ETMC) with $N_f=2+1+1$ flavours of Wilson twisted mass
quarks and Iwasaki gauge action~\cite{Baron:2010bv,Baron:2011sf}. The
details are described in ref.~\cite{Baron:2010bv} and the ensembles
used in this investigation
are summarised in table~\ref{tab:setup}. The twisted mass Dirac
operator in the light -- i.e. up/down -- sector
reads~\cite{Frezzotti:2000nk}
\begin{equation}
\label{eq:Dud}
D_\ell = D_W + m_0 + i \mu_\ell \gamma_5\tau^3
\end{equation}
and in the strange/charm sector~\cite{Frezzotti:2003xj}
\begin{equation}
\label{eq:Dsc}
D_h = D_W + m_0 + i \mu_\sigma \gamma_5\tau^1 + \mu_\delta \tau^3\, ,
\end{equation}
where $D_W$ is the Wilson Dirac operator. The value of $m_0$ was tuned
to its critical value as discussed in
refs.~\cite{Chiarappa:2006ae,Baron:2010bv} in
order to realise automatic $\mathcal{O}(a)$ improvement at maximal
twist~\cite{Frezzotti:2003ni}. Note that the bare twisted masses
$\mu_{\sigma,\delta}$ are related to the bare strange and charm quark
masses via the relation
\begin{equation}
\label{eq:msc}
m_{c,s} = \mu_\sigma\ \pm\ (Z_\mathrm{P}/Z_\mathrm{S})\ \mu_\delta
\end{equation}
with pseudo-scalar and scalar renormalisation constants $Z_\mathrm{P}$
and $Z_\mathrm{S}$. Quark fields in the twisted basis are denoted by
$\chi_{\ell,h}$ and in the physical basis by
$\psi_{\ell,h}$. They are related via the axial rotations
\begin{equation}
\chi_\ell = e^{i\pi\gamma_5\tau^3/4}\psi_\ell\,,\quad\bar\chi_\ell
= \bar\psi_\ell\ e^{i\pi\gamma_5\tau^3/4}\,,\qquad
\chi_h = e^{i\pi\gamma_5\tau^1/4}\psi_h\,,\quad\bar\chi_h
= \bar\psi_h\ e^{i\pi\gamma_5\tau^1/4}\,.
\end{equation}
With automatic $\mathcal{O}(a)$ improvement being the biggest
advantage of tmLQCD at maximal twist, the downside is that flavour
symmetry is broken at finite values of the lattice spacing. This was
shown to affect mainly the mass value of the neutral pion
mass~\cite{Urbach:2007rt,Dimopoulos:2009qv,Baron:2009wt}, however, in
the case of $N_f=2+1+1$ dynamical quarks, it implies the complication of
mixing between strange and charm quarks.
\section{Flavour Singlet Pseudo-Scalar Mesons in $N_f=2+1+1$ tmLQCD}
In order to compute masses of pseudo-scalar flavour singlet mesons we
have to include light, strange and charm contributions to build the
appropriate correlation functions. In the light sector, one appropriate
operator is given by~\cite{Jansen:2008wv}
\begin{equation}
\label{eq:ll}
\frac{1}{\sqrt{2}}(\psi_u i \gamma_5\psi_u + \psi_d i \gamma_5\psi_d)
\quad \to\quad \frac{1}{\sqrt{2}}(\bar\chi_u \chi_u - \bar\chi_d \chi_d)\equiv\ell\, .
\end{equation}
In the strange and charm sector, the corresponding operator reads
\begin{equation}
\label{eq:physsinglet}
\begin{pmatrix}
\bar \psi_c \\
\bar \psi_s \\
\end{pmatrix}
i \gamma_5 \frac{1 \pm \tau^3}{2}
\begin{pmatrix}
\psi_c \\
\psi_s \\
\end{pmatrix}
\quad\to\quad
\begin{pmatrix}
\bar \chi_c \\
\bar \chi_s \\
\end{pmatrix}
\frac{-\tau^1 \pm i \gamma_5\tau^3}{2}
\begin{pmatrix}
\chi_c \\
\chi_s \\
\end{pmatrix}\, .
\end{equation}
In practice we need to compute correlation functions of the
following interpolating operators
\begin{equation}
\label{eq:psequ}
\begin{split}
P_{ss}\ \equiv\ (\bar\psi_s i \gamma_5\psi_s) &=
(\bar\chi_c i \gamma_5\chi_c - \bar\chi_s i \gamma_5\chi_s)/2 -
\frac{Z_\mathrm{S}}{Z_\mathrm{P}}(\bar\chi_s\chi_c +
\bar\chi_c\chi_s)/2\, , \\
P_{cc}\ \equiv\ (\bar\psi_c i \gamma_5\psi_c) &=
(\bar\chi_s i \gamma_5\chi_s - \bar\chi_c i \gamma_5\chi_c)/2 -
\frac{Z_\mathrm{S}}{Z_\mathrm{P}}(\bar\chi_s\chi_c +
\bar\chi_c\chi_s)/2\, . \\
\end{split}
\end{equation}
Note that the sum of pseudo-scalar and scalar contributions
appears with the ratio of renormalisation factors $Z\equiv
Z_\mathrm{S}/Z_\mathrm{P}$, which needs to be
taken into account properly. $Z$ has not yet been determined for
all values of $\beta$ non-perturbatively.
However, for the mass determination, we can avoid this complication
by changing the basis and compute the real and positive
definite correlation matrix
\begin{equation}
\label{eq:corrmatrix}
\mathcal{C} =
\begin{pmatrix}
\eta_{\ell\ell} & \eta_{\ell P_h} & \eta_{\ell S_h} \\
\eta_{P_h \ell} & \eta_{P_h P_h} & \eta_{P_h S_h} \\
\eta_{S_h \ell} & \eta_{S_h P_h} & \eta_{S_h S_h} \\
\end{pmatrix}\, ,
\end{equation}
with the notation
\begin{equation}
P_h \equiv (\bar\chi_c i \gamma_5\chi_c - \bar\chi_s i
\gamma_5\chi_s)/2\, ,\qquad
S_h \equiv (\bar\chi_s\chi_c + \bar\chi_c\chi_s)/2
\end{equation}
and $\eta_{XY}$ denoting the corresponding correlation
function. Masses can determined by solving the generalised eigenvalue
problem~\cite{Michael:1982gb,Luscher:1990ck}
\begin{equation}
\mathcal{C}(t)\ \eta^{(n)}(t,t_0) = \lambda^{(n)}(t, t_0)\ \mathcal{C}(t_0)\
\eta^{(n)}(t,t_0) \, .
\end{equation}
Taking into account the periodic boundary conditions for a meson, we
can determine the effective masses by solving
\begin{equation}
\frac{\lambda^{(n)}(t,t_0)}{\lambda^{(n)}(t+1,t_0)} = \frac{e^{-m^{(n)}
t}+ e^{-m^{(n)}(T-t)}}
{e^{-m^{(n)}(t+1)}+ e^{-m^{(n)}(T-(t+1))}}
\end{equation}
for $m^{(n)}$, where $n$ counts the eigenvalues. The state with the
lowest mass should correspond to the $\eta$ and the second state
to the $\eta'$ meson.
From the components $\eta^{(n)}_{0,1,2}$ of the eigenvectors, we can
reconstruct the physical flavour contents $c^{(n)}_{\ell,s,c}$ from
\begin{equation}
\begin{split}
c_\ell^{(n)} &= \frac{1}{\mathcal{N}^{(n)}} (\eta^{(n)}_0)\\
c_s^{(n)} &= \frac{1}{\mathcal{N}^{(n)}} (- Z \eta^{(n)}_1 +
\eta^{(n)}_2)/\sqrt{2}\\
c_c^{(n)} &= \frac{1}{\mathcal{N}^{(n)}} (- Z \eta^{(n)}_1 -
\eta^{(n)}_2)/\sqrt{2}\\
\end{split}
\end{equation}
with normalisation
\[
\mathcal{N}^{(n)} = \sqrt{(\eta^{(n)}_0)^2 + (Z \eta^{(n)}_1)^2 + (\eta^{(n)}_2)^2}\ .
\]
At this point the ratio $Z\equiv Z_\mathrm{S}/Z_\mathrm{P}$ is needed
again. Assuming for a moment that charm does not contribute
significantly to the $\eta$ and $\eta'$ states, one can extract the
$\eta$-$\eta'$ mixing angle $\phi$ from
\begin{equation}
\label{eq:angle}
\cos(\phi) = c_\ell^{(0)} \approx c_s^{(1)},\quad sin(\phi) =
-c_s^{(0)} = c_\ell^{(1)}
\end{equation}
with $^{(0)}$ ($^{(1)}$) denoting the $\eta$ ($\eta'$) state.
\section{Results}
We have computed all contractions needed for building the correlation
matrix of eq.~(\ref{eq:corrmatrix}). For the connected contributions, we
used stochastic time-slice sources (the so called
``one-end-trick''~\cite{Boucaud:2008xu}). For the disconnected
contributions, we used stochastic volume sources with complex
Gaussian noise~\cite{Boucaud:2008xu}. As discussed in
ref.~\cite{Jansen:2008wv} one can estimate the light disconnected
contributions very efficiently using the identity
\[
D_u^{-1} - D_d^{-1} = -2i\mu_\ell D_d^{-1}\ \gamma_5\ D_u^{-1}\ .
\]
For the heavy sector such a simple relation does not exist, but we
can use the so called hopping parameter variance reduction, which
relies on the same equality as in the mass degenerate two flavour
case (see ref.~\cite{Boucaud:2008xu} and references therein)
\[
D_h^{-1} = B - BHB + B(HB)^2 - B(HB)^3 + D_h^{-1} (HB)^4
\]
with $D_h = (1 + HB) A$, $B = 1/A$ and $H$ the two flavour hopping
matrix. We use $24$ stochastic volume sources per gauge
configuration in both the heavy and the light sector.
We use both local and fuzzed sources to enlarge our correlation
matrix by a factor two. In addition to the interpolating operator
quoted in eqs.~(\ref{eq:ll}) and (\ref{eq:physsinglet}), we also
plan to consider the $\gamma$-matrix combination $i\gamma_0\gamma_5$, which
will increase the correlation matrix by another factor of two. The number of
gauge configurations investigated per ensemble is in most cases
around $1200$, and for ensemble B25.32 is $1500$. Statistical errors
are computed using the bootstrap method with $1000$ samples.
\begin{figure}[t]
\centering
\subfigure[]{\includegraphics[width=.48\linewidth]
{plots/B25.32.M_i_vs_t_3x3.eps}}\quad
\subfigure[]{\includegraphics[width=.48\linewidth]
{plots/flavor_content_3x3_state_1.eps}}
\caption{(a) Effective masses in lattice units determined from solving
the generalised eigenvalue problem with $t_0/a=1$ for ensemble
B25.32. We show the results extracted from a $3\times3$
matrix. (b) squared flavour content of $\eta$ for B25.32.}
\label{fig:effmasses}
\end{figure}
In figure~\ref{fig:effmasses} we show the effective masses determined
from solving the generalised eigenvalue problem for ensemble
B25.32 from a $3\times3$ matrix with local operators only. We kept
$t_0/a=1$ fixed. One observes that the ground state is very well
determined and it
can be extracted from a plateau fit. The second state, i.e. the
$\eta'$, is much more noisy and a mass determination is questionable,
at least from a $3\times3$ matrix. Enlarging the matrix size
significantly reduces the contributions of excited states to the
lowest states and, due to smaller statistical errors at smaller $t$
values, a determination becomes possible. The third state appears to be
in the region where one would expect the $\eta_c$ mass value, however,
the signal is lost at $t/a=5$ already, which makes a reliable
determination not feasible.
\begin{figure}[t]
\centering
\includegraphics[width=.7\linewidth]%
{plots/M_i_vs_M_pi_squared_cosh_fit.eps}
\caption{Preliminary values for $m_\eta$ and $m_{\eta'}$ for two
$\beta$-values in physical units
as a function of the squared pion mass. Filled symbols represent
$m_\eta$, open ones $m_{\eta'}$.}
\label{fig:masssummary}
\end{figure}
In figure~\ref{fig:masssummary} we show the masses of the $\eta$ and
$\eta'$ mesons for the various ensembles we used as a function of the
squared pion mass. In addition we show the corresponding physical
values. The scale was set from $f_\pi$ and $m_\pi$ using the results of
ref.~\cite{Baron:2010bv}. It is clear that the $\eta$ meson
mass can be extracted with high precision, while the $\eta'$ meson
mass requires a larger correlation matrix, which is work in
progress. The comparison with the corresponding physical values
seems to point towards good agreement.
We also determine the flavour content of the two states as explained
above. It turns out that the $\eta$ has a dominant strange quark
content (see right panel of figure~\ref{fig:effmasses}), while the
$\eta'$ is dominated by light quarks. For both the
charm contribution is rather small, however, for the $\eta$ it turns
out to be significantly non-zero. A preliminary determination of the
mixing angle eq.~(\ref{eq:angle}) yields a very stable value of about
$60\,^{\circ}$.
Note that this is the mixing angle to the flavour
eigenstates. The computation of the mixing angle with respect to the
$\eta_0$ and $\eta_8$ states is in progress.
\subsection{Bare Strange and Charm Quark Mass Dependence}
The results displayed in figure~\ref{fig:masssummary} have been
obtained using the bare values of $\mu_\sigma$ and $\mu_\delta$ as
used for the production of the ensembles. Those values, however, did
not lead to the correct values of, for e.g., the kaon and $D$-meson masses,
see e.g. ref.~\cite{Baron:2011sf}. Moreover, the physical strange
and charm quark mass values differ between the $A$ and $B$
ensembles. Hence, figure~\ref{fig:masssummary} is not yet conclusive
with regards to the size of lattice artifacts and extrapolation to the
physical point. What we can learn is that the light quark mass
dependence in both states appears to be rather weak.
We also have two ensembles A80.24 and A80.24s with different bare
values for $\mu_{\sigma,\delta}$ and a retuned value for $\kappa$
but identical parameters
otherwise. Ensemble A80.24s is significantly better
tuned with respect to the physical kaon mass
value~\cite{Baron:2011sf}. As seen in figure~\ref{fig:masssummary},
our results seem to indicate that this change in the bare parameters
also has a significant impact on the $\eta$ meson mass value, while the
$\eta'$ mass is unaffected within the (rather large) errors.
\section{Summary and Outlook}
We presented a computation of $\eta$ and $\eta'$ meson masses from
$N_f=2+1+1$ Wilson twisted mass lattice QCD. The results we obtained
so far are rather encouraging, the $\eta$ meson mass can be determined
with high precision. Also for the $\eta'$ meson mass, we hope to be
able to give more precise results by increasing the correlation matrix
under investigation.
As it is not so easy to tune bare strange and charm quark masses
exactly, we shall in the future use a mixed action approach for strange
and charm quarks. This will not only avoid the complication induced by
flavour symmetry breaking in the twisted mass formulation, but it will
also allow to use the efficient noise reduction techniques in the
heavy sector. However, this method requires a careful investigation of
unitarity breaking effects.
\subsection*{Acknowledgments}
We thank J.~Daldrop, K.~Jansen, C.~McNeile, M.~Petschlies,
M.~Wagner and F.~Zimmermann for useful discussions. We thank the
members of ETMC for the most enjoyable collaboration. The computer
time for this project was made available to us by the John von
Neumann-Institute for Computing (NIC) on the JUDGE and Jugene systems
in J{\"u}lich. This project was funded by the DFG as a
project in the SFB/TR 16. The open source software packages
tmLQCD~\cite{Jansen:2009xp} and Lemon~\cite{Deuzeman:2011wz} have been
used.
\bibliographystyle{h-physrev5}
|
\section{Introduction}
\label{IntroSection}
When the technique of integration by parts in $d$ dimensions was first proposed by Tkachov and Chetyrkin~\cite{Chetyrkin}, it represented a major breakthrough in the study of perturbative gauge theories at the multi-loop level.\footnote{For the reader less familiar with integration by parts, we strongly recommend Smirnov's excellent book on Feynman integral calculus~\cite{EFInts}. He explains the technique and works through a number of simple (and non-simple) examples.} Their discovery was of fundamental practical importance, as it allowed researchers to perform many multi-loop calculations that were previously thought to be intractable. Furthermore, the idea of the technique is simple to describe. By taking integrals of derivatives in $d$ dimensions, one generates a tower of equations for the Feynman integrals belonging to a particular topology. Then one tries to solve these equations, either by inspection or via some systematic procedure, for an independent basis of master integrals.
Unfortunately, the total number of equations generated in this way grows rapidly with the number of loops and external states in the integral topology under consideration. As a consequence, the solution of the so-called integration by parts relations is complicated for all but the simplest examples. For many years after the technique was introduced, no systematic procedure for the solution of integration by parts relations was known and it was therefore only possible to apply the method to simple integral topologies.\footnote{Of course, simple is a relative term. For a rather impressive early application of the method at the three-loop level see the long write-up for {\sc MINCER}~\cite{MINCER}.} The situation changed just over a decade ago with the introduction of a Gaussian elimination-like solution algorithm. This algorithm, due to Laporta~\cite{Laporta}, was a crucial step forward because it allowed researchers to apply the integration by parts technique to highly non-trivial problems for which an {\it ad hoc} approach would be impractical if not impossible. Although Laporta's algorithm has been tremendously successful, it has long been known that it may lead to master integrals with doubled propagator denominators.
To be more precise, for an $L$-loop topology, let $V = \{\ell_1,\dots,\ell_L,k_1\dots,k_{N}\}$ be the set of loop momenta together with a set of $N$ independent external momenta and let $v$ be a generic element of this set. Normally, one generates the set of integration by parts relations in an obvious way, considering each $v \in V$ in turn and writing all possible equations of the form
\bea
0 = \prod_{i = 1}^L\left( \int {d^d \ell_i\over (2 \pi)^d}\right)~ {\partial\over \partial \ell_j}\cdot\left( {v^{(j)}~ \mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}\over D_1(\ell_1,\dots\ell_L)^{b_{1}}\cdots D_m(\ell_1,\dots\ell_L)^{b_{m}}} \right)\,.
\label{IBP1}
\eea
Crucially, the indices $a_i$ and $b_i$ satisfy certain boundary conditions; typically the irreducible numerators (here we assume that each $\mathcal{N}_i$ has mass dimension two) for a given topology enter raised to, at most, some relatively small positive integer and, in some cases, certain propagator denominators are constrained to have non-negative indices since otherwise the resulting integrals vanish in dimensional regularization. This set of equations together with the appropriate boundary conditions can then be fed into Laporta's algorithm. The price one pays for being able to trivially generate the set of integration by parts relations in this fashion is twofold. Not only will the algorithm (Laporta or typical variation thereof~\cite{AIR,FIRE,REDUZE}) used to solve the resulting system of equations typically have to eliminate an enormous number of spurious integrals as it attempts to solve the system, it is not straightforward to ensure that each integral in the basis ultimately returned by the algorithm has the property that $b_i = 1$.
Although there is nothing wrong with integrals that have some $b_i > 1$, they may be inconvenient for particular multi-loop applications. For example, in computational approaches based on generalized unitarity, one would like to have a basis with well-defined unitarity cuts in all channels (see {\it e.g.}~\cite{KLunitarity} and~\cite{MOunitarity} for recent work in this direction at two loops). It is unclear how one would make sense of integrals with doubled propagator denominators in such a framework. Gluza, Kajda, and Kosower (hereafter referred to as GKK) also argued in~\cite{GKK} that master integrals with doubled propagator denominators can be significantly harder to expand in $\epsilon$ than those without.
The idea of the GKK procedure is relatively easy to state now that the stage is set. With the above motivation, GKK found that they could completely avoid the introduction of doubled propagators by imposing $b_i = 1$ from the beginning. They observed that, generically, one expects doubled propagators for the simple reason that the derivatives in eqs. (\ref{IBP1}) act on the propagator denominators, $D_k$. They also recognized that there is no good reason why one ought to consider the elements of $V$ one at a time; one can generalize eqs. (\ref{IBP1}) by replacing $v^{(j)}$ with a general linear combination of the elements of $V$:
\bea
0 = \prod_{i = 1}^L\left( \int {d^d \ell_i\over (2 \pi)^d}\right)~ {\partial\over \partial \ell_j}\cdot\left( {\sum_{i = 1}^{L + N}\alpha_i^{(j)} v_i^{(j)}~ \mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}\over D_1(\ell_1,\dots\ell_L)^{b_{1}}\cdots D_m(\ell_1,\dots\ell_L)^{b_{m}}} \right)\,.
\eea
It is also convenient at this point to combine together some of the equations by summing over $j$:
\begin{changemargin}{-.2 in}{0 in}
\bea
0 = \prod_{i = 1}^L\left( \int {d^d \ell_i\over (2 \pi)^d}\right)~\sum_{j=1}^L {\partial\over \partial \ell_j}\cdot\left( {\sum_{i = 1}^{L + N}\alpha_i^{(j)} v_i^{(j)}~ \mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}\over D_1(\ell_1,\dots\ell_L)^{b_{1}}\cdots D_m(\ell_1,\dots\ell_L)^{b_{m}}} \right)\,.
\eea
\end{changemargin}
In a nutshell, the GKK strategy is to start with all the $b_i$ equal to one
\bea
0 = \prod_{i = 1}^L\left( \int {d^d \ell_i\over (2 \pi)^d}\right)~\sum_{j=1}^L {\partial\over \partial \ell_j}\cdot\left( {\sum_{i = 1}^{L + N}\alpha_i^{(j)} v_i^{(j)}~ \mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}\over D_1(\ell_1,\dots\ell_L)\cdots D_m(\ell_1,\dots\ell_L)} \right)\,
\label{IBP2}
\eea
and then choose the coefficients $\alpha_i^{(j)}$ in such a way that the numerator exactly cancels all unwanted, derivative-generated powers of the propagator denominators. In other words, for each $k$, we demand that
\be
\sum_{j = 1}^L \sum_{i = 1}^{L + N} \alpha_i^{(j)}\,{\partial D_k \over \partial \ell_j}\cdot v_i^{(j)}~ \mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q} = - 2 T^{(k,a_1,\dots,a_q)} D_k \,,
\label{IBP3}
\ee
where $T^{(k,a_1,\dots,a_q)}$ is some polynomial built out of the irreducible numerators and Lorentz invariant combinations of the vectors in $V$.\footnote{We write $- 2 T^{(k,a_1,\dots,a_q)}$ instead of $T^{(k,a_1,\dots,a_q)}$ so that we will ultimately arrive at the same form that GKK did in their eqs. (4.1).} Each independent set of $L (L + N)$ coefficients $\alpha_i^{(j)}$ satisfying eqs. (\ref{IBP3}), upon substitution into eqs. (\ref{IBP2}), yields a unitarity-compatible integration by parts relation, by construction free of doubled propagators.
The downside of this novel approach is that now one needs to find a way to generate complete sets of $\alpha_i^{(j)}$ coefficients and it turns out that this is not a straightforward task. Although GKK did propose a solution\footnote{Actually, GKK presented two distinct algorithms for the generation of complete sets of unitarity-compatible integration by parts relations. In what follows, when we refer to GKK's ``solution'' or ``solution algorithm'' with no additional qualifier, we are referring to their best solution (what they call Algorithm III).} to this problem in~\cite{GKK}, they admitted that their solution was somewhat provisional and that there is likely room for improvement. The GKK solution relies heavily on the use of Gr\"{o}bner bases, important constructs in computational commutative algebra which are, however, notoriously difficult to compute in practice~\cite{alggeom}.\footnote{For the reader less well-versed in computational commutative algebra, we can strongly recommend the very well-written and concise survey by Adams and Loustaunau~\cite{AL}. Most of the relevant mathematical concepts are also defined and briefly explained in the GKK paper~\cite{GKK}.} In this paper we propose a completely different solution to the problem of generating complete sets of unitarity-compatible integration by parts relations. As we shall see, our Algorithm \ref{alg} is based entirely on simple linear algebra and, in particular, completely avoids the use of Gr\"{o}bner bases.
This article is organized as follows. In section~\ref{review} we describe more precisely the problem we wish to solve and introduce some useful notation. In section~\ref{body} we present the main result of this paper, Algorithm \ref{alg}, and talk the reader through it. In Section \ref{example}, we give a detailed example of how our algorithm works in practice. In section~\ref{ConclusionSection} we present our conclusions and outline our plan for future research.
\section{Preliminaries}
\label{review}
In this section we take a closer look at the reduction procedure introduced by GKK and discuss its strengths and weaknesses. Our initial goal will be to precisely set up the mathematical problem that lies at the heart of the GKK procedure and discuss why (in the opinion of both GKK and the present author) the solution presented in~\cite{GKK} is not likely to be the best one possible. We will then explain our approach to the problem and illustrate with a very simple example what precisely Algorithm \ref{alg} is designed to do.
As we left them, eqs. (\ref{IBP3}) look rather cumbersome. We can clean them up by absorbing the numerator polynomial $\mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}$ into the $\alpha_i^{(j)}$
\be
\sum_{j = 1}^L\sum_{i = 1}^{L + N} \alpha_i^{(j)}\,{1 \over 2}{\partial D_k \over \partial \ell_j}\cdot v_i^{(j)} + T^{(k, \alpha)} D_k = 0 \,,
\label{IBP4}
\ee
with the understanding that now the $\alpha_i^{(j)}$ are dimensionful. We will have to take this into account in our search for independent solutions. To make further progress, GKK expressed (\ref{IBP4}) as a matrix equation:
\be
\beta \cdot E = {\bf 0}\,,
\label{IBP5}
\ee
where
\be
\beta = \Big(\alpha_1^{(1)},\dots,\alpha_{L+N}^{(1)},\dots\dots,\alpha_1^{(L)},\dots,\alpha_{L+N}^{(L)},T^{(1, \alpha)},\dots,T^{(m, \alpha)}\Big)\,.
\ee
Given an explicit expression for $V$ and an ordering on the set of propagator denominators, the entries of $E$ can be straightforwardly read off from (\ref{IBP4}). For an explicit example, we refer the reader to eq. (5.3) of~\cite{GKK}. In eq. (5.3), GKK wrote out $E$ explicitly for the planar massless double box, choosing $\{\ell_1^2, \ell_1^2 - 2\ell_1\cdot k_1, \ell_1^2 - 2\ell_1\cdot k_1 -2 \ell_1\cdot k_2 +s_{1 2}, \ell_2^2, \ell_2^2 - 2 \ell_2\cdot k_4, \ell_2^2 - 2 \ell_2\cdot k_3 - 2 \ell_2\cdot k_4 + s_{12},\ell_1^2 + \ell_2^2 + 2 \ell_1\cdot \ell_2\}$ for the ordering on the set of propagator denominators (this ordering fixes the sequence of the columns of $E$).\footnote{We should also point out that the precise definitions GKK made for the $\alpha_i^{(j)}$ (what they call $c_j^{(i)}$) and $T^{(k,\alpha)}$ (what they call $u_k$) do not seem to exactly reproduce eq. (5.3). However, the differences can be taken into account by appropriately rescaling the unknowns and are therefore unimportant.} The point of making this rearrangement is that now the problem of determining all independent sets of $\alpha_i^{(j)}$ coefficients looks like a well-known, well-studied problem from computational commutative algebra: the computation of the syzygy module\footnote{Given a set of generators for a module $\mathcal{M}$, $\{{\bf M}_1,\dots,{\bf M}_r\}$, the syzygy module of $\mathcal{M}$ is simply the set of all $\beta = (b_1,\dots,b_r)$ such that $\sum_{i = 1}^r b_i \cdot {\bf M}_i = {\bf 0}$. In this paper, we will deal only with submodules of $F[x_1,\dots,x_n]^r$ and $F[x_1,\dots,x_n]^m$ for a field, $F$.} of a module for which one has an explicit set of generators.
Actually, as explained by GKK, it suffices to solve this problem for ideals, since given a set of generators for a module $\mathcal{M}$, one can easily define a set of generators for an ideal with exactly the same syzygies. Suppose the set $\{{\bf M}_1,\dots,{\bf M}_r\}$ generates $\mathcal{M}$. By simply taking the dot product of each ${\bf M}_i$ with a tuple of dummy variables satisfying the relations $t_i t_j = 0$, $\{t_1,\dots,t_m\}$, we can convert our generating set for $\mathcal{M}$ into a generating set for an ideal $\mathcal{I}$ canonically associated to $\mathcal{M}$. The class of ideals canonically associated to the modules described by eqs. (\ref{IBP4}) can be taken to be homogeneous of uniform degree two. We can see this as follows. By definition, each parameter entering into the matrix $E$ of eq. (\ref{IBP5}) is of the form ${\partial D_k \over \partial \ell_j}\cdot v_i^{(j)}$ or $D_k$ and is going to have mass dimension two. Therefore, it makes sense to take every independent Lorentz product that can arise in $E$ to be an independent variable and the class of modules described by eqs. (\ref{IBP4}) to be homogeneous of uniform degree one. It then follows, after applying the canonical map described above, that the ideals of interest are homogeneous of uniform degree two. In the end, we find that each generator of $\mathcal{I}$ has terms of the schematic form $a\, t_j\,x_k$, $a \in F$ for some field $F$.\footnote{In this paper, $F$ is given by the field of rational functions of the auxiliary parameters in the matrix $E$ ({\it e.g.} $\chi_{14}$ in the case of the planar massless double box treated in detail by GKK) of eq. (\ref{IBP5}). However, for practical purposes, it is more convenient to simply assign prime numbers to the dimensionless auxiliary parameters and work over $\mathbb{Q}$.}
Before continuing, we need to introduce a little more notation and point out an important fact about the syzygies of generating sets for homogeneous ideals of uniform degree. Given an ideal generated by $P = (p_1,\dots,p_r)$, we define $S_d(P)$ to be the set of all syzygies of $P$ such that $\beta = (b_1,\dots,b_r) \in S_d(P)$ implies that each $b_i$ is a polynomial of uniform degree $d$. We will typically refer to $S_d(P)$ as the set of all degree $d$ syzygies of $P$. It turns out that, for the case of homogeneous ideals of uniform degree, the syzygy module, $S(P)$, is a graded module~\cite{LASyz}:
\be
S(P) = \bigoplus_{d = 0}^\infty S_d(P)\,.
\ee
This means that, for homogeneous ideals of uniform degree, it suffices to search for degree $d$ syzygies.
Determining a basis for the syzygy module of a generic ideal is known to be a very difficult problem~\cite{alggeom} (and is very much an active area of mathematical research). Therefore, one needs a dedicated solution algorithm, tailor-made for the class of ideals described above. As mentioned in the introduction, the solution algorithm employed by GKK relies heavily on the use of Gr\"{o}bner bases. GKK chose the well-known Buchberger algorithm~\cite{Buchberger} to compute their Gr\"{o}bner bases. In their paper, GKK pointed out that there have been a number of recent attempts to improve on Buchberger's algorithm (see {\it e.g.}~\cite{Eder} for a description of one of the most promising of these recent attempts, based on Faug\`{e}re's $F_5$ algorithm~\cite{Faugere}) and concluded that their Buchberger-based approach was not likely to be optimal. In this paper we rethink their approach at a more fundamental level.
Certainly, one could attempt to compute complete sets of syzygies using an approach based on Faug\`{e}re's $F_5$ algorithm~\cite{SyzF5} or some other improved algorithm for the computation of Gr\"{o}bner bases. However, it is actually no longer clear that one should use Gr\"{o}bner bases at all. Quite recently (after the appearance of~\cite{GKK}) it was shown in~\cite{LASyz} that, remarkably, bases for the modules of syzygies of special classes of ideals can be computed using simple linear algebra. Actually, at this stage, the reader may be wondering what stopped GKK from using a linear algebra-based approach in the first place. Na\"{i}vely, linear algebra seems to offer a very easy way to compute syzygies.
To illustrate why the ideas of reference~\cite{LASyz} are non-trivial, we consider the ideal generated by $P = \{x_1 - x_2, 2 x_2 - x_1\}$ and attempt to compute its syzygies using linear algebra.\footnote{For the sake of definiteness, suppose that we are working with polynomials in $\mathbb{Q}[x_1,x_2]$.} By inspection, we see that $P$ has no degree zero syzygies. By definition, a degree one syzygy of $P$ must have the form $(c_1 x_1 + c_2 x_2, c_3 x_1 + c_4 x_2)$ for some elements $(c_1,c_2,c_3,c_4)$ of $\mathbb{Q}$. Starting with this ansatz, we can take the dot product with $P$
\be
(c_1 x_1 + c_2 x_2, c_3 x_1 + c_4 x_2)\cdot P = (c_1 - c_3) x_1^2 + (-c_1 + c_2 + 2 c_3 - c_4) x_1 x_2 + (-c_2 + 2 c_4) x_2^2\,,
\ee
set the coefficient of each monomial in the sum to zero, and solve the resulting system of equations. It turns out that there is a one-parameter family of solutions which we parametrize by $c_4$:
\be
(c_1,c_2,c_3,c_4) = (-c_4, 2 c_4, -c_4, c_4)\,.
\ee
We can arbitrarily set $c_4 = 1$ to obtain a basis for $S_1(P)$, $(- x_1 + 2 x_2, - x_1 + x_2)$. So far so good. By definition, a degree two syzygy of $P$ must have the form $(c_1 x_1^2 + c_2 x_1 x_2 + c_3 x_2^2, c_4 x_1^2 + c_5 x_1 x_2 + c_6 x_2^2)$. Going through the same procedure, we arrive at a two-parameter family of solutions which we parametrize by $c_5$ and $c_6$:
\be
(c_1,c_2,c_3,c_4,c_5,c_6) = (- c_5 - c_6, 2 c_5 + c_6, 2 c_6, - c_5 - c_6, c_5, c_6)\,.
\ee
By replacing $(c_5,c_6)$ with each of the standard basis vectors for $\mathbb{R}^2$ in turn, we find that $(-1, 2, 0, -1, 1, 0)$ and $(-1, 1, 2, -1, 0, 1)$ form a basis for the set of solutions. In order to map these solutions back to degree two syzygies of $P$, we partition each vector into two non-overlapping subsets of length three without changing the overall ordering of the entries
\be
\Big\{\Big(\{-1,2,0\},\{-1,1,0\}\Big), \Big(\{-1,1,2\},\{-1,0,1\}\Big)\Big\}
\ee
and then dot each resulting 3-tuple into $\{x_1^2, x_1 x_2, x_2^2\}$:
\bea
&&\Big\{\Big(\{x_1^2, x_1 x_2, x_2^2\}\cdot\{-1,2,0\},\{x_1^2, x_1 x_2, x_2^2\}\cdot\{-1,1,0\}\Big),
\nonumber \\ &&\Big(\{x_1^2, x_1 x_2, x_2^2\}\cdot\{-1,1,2\},\{x_1^2, x_1 x_2, x_2^2\}\cdot\{-1,0,1\}\Big)\Big\}
\nonumber \\ &&= \Big\{\Big(- x_1^2 + 2 x_1 x_2, - x_1^2 + x_1 x_2\Big), \Big(- x_1^2 + x_1 x_2 + 2 x_2^2, - x_1^2 + x_2^2\Big)\Big\}\,.
\label{examp1}
\eea
This overly formal description of the mapping back to syzygies of $P$ is overkill for the example at hand but will be useful later on.
We might be tempted to conclude that we have found two new, linearly independent, degree two syzygies of $P$. This, however, is not true. It turns out that the syzygies of eq. (\ref{examp1}) can be expressed as multiples of $(- x_1 + 2 x_2, - x_1 + x_2)$. Explicitly, we have
\be
\Big(- x_1^2 + 2 x_1 x_2, - x_1^2 + x_1 x_2\Big) = x_1 (- x_1 + 2 x_2, - x_1 + x_2)
\ee
and
\be
\Big(- x_1^2 + x_1 x_2 + 2 x_2^2, - x_1^2 + x_2^2\Big) = (x_1 + x_2)(- x_1 + 2 x_2, - x_1 + x_2)\,.
\ee
Besides highlighting the profound difference between linear independence in a vector space and linear independence in a module, this example shows what goes wrong if one tries to compute the syzygies of homogeneous ideals of uniform degree using na\"{i}ve linear algebra. One is able to compute syzygies in a straightforward manner but not a basis of linearly independent syzygies. It is therefore remarkable that, under appropriate assumptions, this obstacle is actually surmountable. In the next section, drawing upon some of the ideas introduced in~\cite{LASyz}, we present a simpler linear algebra-based alternative to the solution adopted by GKK.
\section{The Algorithm}
\label{body}
The purpose of this section is to give an explicit pseudo-code detailing our solution to the problem defined in Section \ref{review} and to thoroughly comment it. The pseudo-code we present (Algorithm \ref{alg} and associated subroutines) is quite explicit and should allow the reader to fashion a crude implementation of our algorithm in {\tt Maple} or {\tt Mathematica} with very little effort (beyond that required to understand the algorithm in the first place). We should emphasize that we do not in any way claim that an implementation based on our pseudo-code is an {\it effective} implementation. The pseudo-code given below is intended to be maximally clear as opposed to maximally efficient.\footnote{For example, the first statement in the upper branch of the If statement in Algorithm \ref{alg} is completely superfluous and was included only because, in our opinion, it makes the functionality of Subroutine \ref{sub1} significantly easier to understand.} Before discussing the non-trivial features of Algorithm \ref{alg}, we need to make a few more definitions.
First, let $M_d^{(n)} = \Big\{X_1^{(d)},\dots,X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)}\Big\}$ be the set of all monomials of degree $d$ built out of the $n$ variables $\{x_1,\dots,x_n\}$.\footnote{Note that there are precisely $\frac{(d+n-1)!}{d!(n-1)!}$ monomials of degree $d$ built out of $n$ variables. This result follows immediately once one recognizes that the monomial counting problem is isomorphic to one of the usual ``ball-box'' counting problems of enumerative combinatorics. We refer the unfamiliar reader to Section 1.4 of Stanley's textbook on the subject~\cite{Stanley}.} For example, $M_2^{(2)} = \{x_1^2, x_1 x_2, x_2^2\}$. For definiteness, we will always order sets of monomials lexicographically. This choice of ordering is just a choice and has no deeper significance. Now, if we have in hand a sequence of $r$ homogeneous polynomials of uniform degree two, $P_0 = \{p_1,\dots,p_r\}$, then $P_d$ is defined to be the outer product of $M_d^{(n)}$ and $P_0$:
\be
P_{d} = \Big\{p_1 X_1^{(d)},\dots,p_1 X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)},\dots\dots,p_r X_1^{(d)},\dots,p_r X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)}\Big\}\,.
\ee
Clearly, $P_d$ is a set of $\frac{r (d+n-1)!}{d!(n-1)!}$ homogeneous polynomials of uniform degree $d + 2$. For example, if we take $P_0 = P = \{x_1 - x_2, 2 x_2 - x_1\}$ it is easy to see that $P_2 = \{(x_1 - x_2)x_1^2, (x_1 - x_2)x_1 x_2, (x_1 - x_2)x_2^2, (2 x_2 - x_1)x_1^2, (2 x_2 - x_1)x_1 x_2, (2 x_2 - x_1)x_2^2\}$. The construction of $P_2$ can be thought of as an intermediate step towards the extraction of the degree two syzygies of $P$. Instead of solving the system $(c_1 x_1^2 + c_2 x_1 x_2 + c_3 x_2^2, c_4 x_1^2 + c_5 x_1 x_2 + c_6 x_2^2)\cdot \{x_1 - x_2, 2 x_2 - x_1\} = 0$, we can construct $P_2$ and then solve the system $(c_1,c_2,c_3,c_4,c_5,c_6)\cdot \{(x_1 - x_2)x_1^2, (x_1 - x_2)x_1 x_2, (x_1 - x_2)x_2^2, (2 x_2 - x_1)x_1^2, (2 x_2 - x_1)x_1 x_2, (2 x_2 - x_1)x_2^2\} = 0$. So, instead of trying to work with $S_d(P_0)$ directly, we can work with the vector space $S_0(P_d)$. This natural correspondence between basis vectors of $S_0(P_d)$ and degree $d$ syzygies of $P_0$ is an essential part of Algorithm \ref{alg}, which is why we took the time to carefully describe it while working through the illustrative example at the end of Section \ref{review}. However, it is important to remember that this map does not necessarily yield independent elements of $S(P_0)$; to work as advertised, our solution algorithm must be able to determine, using only linear algebra, what basis vectors of $S_0(P_d)$ correspond to new syzygies of $P_0$, linearly independent of those already determined.
\newpage
\begin{Algorithm}[H]
\caption{}\label{alg}
\begin{algorithmic}
\REQUIRE A minimal generating set for a homogeneous ideal of uniform degree two, $P_0 = \{p_1(t_1,\dots,t_m,x_1,\dots,x_n),\dots,p_r(t_1,\dots,t_m,x_1,\dots,x_n)\}$. The $p_i$ are further constrained to have terms of the schematic form $a\, t_j\,x_k$ with $a \in F$ and $t_i t_j = 0$ for all $i$ and $j$. ${\rm SyzList}_0 = B_0 = \{{\bf 0}\}$. For a given integral topology with generic irreducible numerator $\mathcal{N}_1(\ell_1,\dots\ell_L)^{a_1}\cdots \mathcal{N}_q(\ell_1,\dots\ell_L)^{a_q}$, $\Delta$ is the largest value of $\sum_{i = 1}^q a_i$ allowed by the boundary conditions on the $a_i$ (we assume that each $\mathcal{N}_i$ has mass dimension two). $M_d^{(n)} = \Big\{X_1^{(d)},\dots,X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)}\Big\}$ is the lexicographically ordered set of all monomials of degree $d$ built out of the $n$ variables $\{x_1,\dots,x_n\}$.
\STATE
\STATE $\bullet$Set $d = 1$
\WHILE{$d \leq \Delta$}
\IF{$B_{d-1} \neq \{{\bf 0}\}$}
\STATE $\bullet$Take the outer product of $P_0$ and $M_{d-1}^{(n)}$ to determine $P_{d-1}$: \\~~~\,$P_{d-1} = \Big\{p_1 X_1^{(d-1)},\dots,p_1 X_{\frac{(d+n-2)!}{(d-1)!(n-1)!}}^{(d-1)},\dots\dots,p_r X_1^{(d-1)},\dots,p_r X_{\frac{(d+n-2)!}{(d-1)!(n-1)!}}^{(d-1)}\Big\}$
\STATE $\bullet$\,Take the outer product of $P_0$ and $M_{d}^{(n)}$ to determine $P_{d}$: \\~~~\,$P_{d} = \Big\{p_1 X_1^{(d)},\dots,p_1 X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)},\dots\dots,p_r X_1^{(d)},\dots,p_r X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)}\Big\}$
\STATE $\bullet$Apply Subroutine \ref{sub1} to each element of $B_{d-1}$ and call the union of the results $A$:
\\~~~\,$A = \bigcup\limits_{i=1}^{|B_{d-1}|} \sigma(\alpha_i)$
\STATE $\bullet$Regard the elements of $A$ as the rows of a matrix, find a row echelon form of this
\\~\, matrix, keep only rows with at least one non-zero entry, and call the result $C$
\STATE $\bullet$Replace with zero the $s$ entries of $P_d$ which correspond
\\~\,to the pivot columns of $C$ and call the result $G$
\STATE $\bullet$Apply Subroutine \ref{sub2} to $G$ and call the result $D$:
\\~~~\,$D = {\rm SSyz}(G)$
\STATE $\bullet$If $D \neq \{{\bf 0}\}$ then $B_d = C ~\bigcup~ D$ and if $D = \{{\bf 0 }\}$ then $B_d = C$
\STATE $\bullet$Apply Subroutine \ref{sub3} to $D$ to determine ${\rm SyzList}_d$:
\\~~~\,${\rm SyzList}_d = {\rm ToSyz}(D)$
\ELSE
\STATE $\bullet$\,Take the outer product of $P_0$ and $M_{d}^{(n)}$ to determine $P_{d}$: \\~~~\,$P_{d} = \Big\{p_1 X_1^{(d)},\dots,p_1 X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)},\dots\dots,p_r X_1^{(d)},\dots,p_r X_{\frac{(d+n-1)!}{d!(n-1)!}}^{(d)}\Big\}$
\STATE $\bullet$Apply Subroutine \ref{sub2} to $P_d$ and call the result $D$:
\\~~~\,$D = {\rm SSyz}(P_d)$
\STATE $\bullet B_d = D$
\STATE $\bullet$Apply Subroutine \ref{sub3} to $D$ to determine ${\rm SyzList}_d$:
\\~~~\,${\rm SyzList}_d = {\rm ToSyz}(D)$
\ENDIF
\STATE $\bullet d = d+1$
\ENDWHILE
\RETURN $\left(\bigcup\limits_{d=1}^\Delta~ {\rm SyzList}_d \right)\Big\backslash \{ {\bf 0}\}$
\end{algorithmic}
\end{Algorithm}
\floatname{algorithm}{Subroutine}
\addtocounter{algorithm}{-1}
\begin{Algorithm}[h!]
\caption{${\rm \sigma}(\alpha)$ maps $\alpha \in B_{d-1}$ to $n$ linearly independent elements of $S_0(P_d)$}\label{sub1}
\begin{algorithmic}
\REQUIRE The objects introduced in Algorithm \ref{alg} and $\alpha = \Big(A_1,\dots,A_{r(d+n-2)!\over (d-1)!(n-1)!}\Big)$
\\\qquad\qquad $= A_1 p_1 X_1^{(d-1)} + \cdots + A_{\frac{r(d+n-2)!}{(d-1)!(n-1)!}} p_r X_{\frac{(d+n-2)!}{(d-1)!(n-1)!}}^{(d-1)} \in B_{d-1}$
\STATE
\STATE
\RETURN $\Big\{A_1 p_1 X_1^{(d-1)}x_1 + \cdots + A_{\frac{r(d+n-2)!}{(d-1)!(n-1)!}} p_r X_{\frac{(d+n-2)!}{(d-1)!(n-1)!}}^{(d-1)} x_1,\dots\dots,A_1 p_1 X_1^{(d-1)} x_n + \cdots$
\\\qquad\qquad$\cdots + A_{\frac{r(d+n-2)!}{(d-1)!(n-1)!}} p_r X_{\frac{(d+n-2)!}{(d-1)!(n-1)!}}^{(d-1)} x_n\Big\}$
\end{algorithmic}
\end{Algorithm}
\begin{Algorithm}[h!]
\caption{${\rm SSyz}(G)$ computes a basis for $S_0(G)$}\label{sub2}
\begin{algorithmic}
\REQUIRE The objects introduced in Algorithm \ref{alg} and a sequence, $G = \{q_1,\dots,q_n\}$, of homogeneous polynomials in the variables $\{t_1,\dots,t_m,x_1,\dots,x_n\}$ with coefficients in $F$.
\STATE
\STATE $\bullet$ Introduce a sequence of $|G|$ dummy variables:
\\~~~\,$Y = \{y_1,\dots,y_{|G|}\}$
\STATE $\bullet$Expand the dot product $G\cdot Y$ and partition the terms in the sum into equivalence classes,
\\~\,declaring terms equivalent if and only if they contain the same monomial in the variables
\\~\,$\{t_1,\dots,t_m,x_1,\dots,x_n\}$
\STATE $\bullet$Set the sum of terms in each equivalence class equal to zero, collect the resulting set of
\\~\,equations constraining the $y_i$, and drop any equations that are obviously redundant
\STATE $\bullet$Put the remaining equations into matrix form:
\\~~~\,$Q~Y = {\bf 0}$
\RETURN a basis for ${\rm Null}(Q)$
\end{algorithmic}
\end{Algorithm}
\begin{Algorithm}[h!]
\caption{${\rm ToSyz}(D)$ maps $\nu$ elements of $S_0(P_d)$ to $\nu$ elements of $S_d(P_0)$}\label{sub3}
\begin{algorithmic}
\REQUIRE The objects introduced in Algorithm \ref{alg} and a list, $D = \{d_1,\dots,d_\nu\}$, $d_i \in S_0(P_d)$.
\STATE
\STATE $\bullet$Partition each vector $d_i$ into $r$ distinct non-overlapping sequences of length ${(d+n-1)!\over d!(n-1)!}$,
\\~\,taking care to preserve the order of the entries of each $d_i$:
\\~~~\,$d_i \rightarrow \left\{z_1^{(i)},\dots,z_r^{(i)}\right\}$
\RETURN $\Big\{\left(M_d^{(n)}\cdot z_1^{(1)},\dots,M_d^{(n)}\cdot z_r^{(1)}\right),\dots\dots,\left(M_d^{(n)}\cdot z_1^{(\nu)},\dots,M_d^{(n)}\cdot z_r^{(\nu)}\right)\Big\}$
\end{algorithmic}
\end{Algorithm}
We now describe how Algorithm \ref{alg} works in some detail. By assumption, the $m$ propagator denominators of the integral topology under consideration form a linearly independent set.\footnote{If this is not the case, one should first reduce to simpler integral topologies satisfying this condition before attempting to apply Algorithm \ref{alg}. GKK explain this procedure in some detail in~\cite{GKK}.} This implies that $P_0$ is a minimal generating set for the ideal under consideration and there are no non-trivial syzygies of degree zero. Since Algorithm \ref{alg} proceeds incrementally in the degree of the syzygies, it is convenient to define two bookkeeping lists, $B_d$ and ${\rm SyzList}_d$, indexed by $d$. $B_d$ is a basis for the vector space $S_0(P_d)$ and ${\rm SyzList}_d$ is a set of linearly independent elements of $S_d(P_0)$ which are, for $d > 1$, also linearly independent of all syzygies of degree $\leq d - 1$ in the set $\bigcup_{i = 1}^{d - 1} {\rm SyzList}_i$ (we will call the elements of ${\rm SyzList}_d$ new degree $d$ syzygies of $P_0$). Due to the fact that $P_0$ has no degree zero syzygies, both lists are initialized to $\{{\bf 0}\}$ and $d$ is initialized to one. Since $B_0 = \{{\bf 0}\}$, Algorithm \ref{alg} always starts with a pass through the lower branch of the If statement. Using Subroutine \ref{sub2}, Algorithm \ref{alg} determines a basis for $S_0(P_1)$ along the lines described in the example near the end of Section \ref{review}. For the sake of discussion, we assume that this basis is non-trivial. This basis, $D$, is then used to determine both $B_1$ and, via Subroutine \ref{sub3}, ${\rm SyzList}_1$. In this section we will not step into Subroutines \ref{sub2} or \ref{sub3}. There is nothing non-trivial about them and we will in any case go through the subroutines once explicitly in Section \ref{example}.
Increment $d$ to two. Subroutine \ref{sub1} lifts the syzygies in $B_1$ to syzygies in $S_0(P_2)$ in every independent way possible. This is accomplished by simply multiplying each of the elements of $B_1$ (written out explicitly in terms of their coordinates, the elements of $P_1$) by each one of the variables $\{x_1,\dots,x_n\}$ in turn. The results can then be interpreted as syzygies in $S_0(P_2)$ (written out explicitly in terms of their coordinates, the elements of $P_2$). Clearly, if $\beta$ is a scalar syzygy of $P_1$, $\beta\cdot (x_i\,P_1) = 0$ for each $1 \leq i \leq n$ by linearity. This implies that $\beta$ can be interpreted as a scalar syzygy of $P_2$ in $n$ independent ways. Given $\nu$ syzygies in $S_0(P_1)$, Subroutine \ref{sub1} produces $\nu\, n$ syzygies in $S_0(P_2)$ (some of which may be linearly dependent). As we shall see, Subroutine \ref{sub1} is a crucial first step towards determining which syzygies in $S_0(P_2)$ correspond to new degree two syzygies of $P_0$ and which do not.\footnote{The reader who has read reference~\cite{LASyz} might be concerned that we have not yet spoken about the principal syzygies of our ideals. Actually, the fact that the $t_i$ are simply coordinate variables with no independent existence of their own (they satisfy $t_i t_j = 0$ for all $i$ and $j$) implies that the ideals of interest to us have no principal syzygies at all.}
Since the syzygies in $S_0(P_2)$ produced by Subroutine \ref{sub1} are not guaranteed to be linearly independent, the next step is to put the output of Subroutine \ref{sub1} into row echelon form, discard all rows without non-zero entries, and call the result $C$. Since, by construction, each of the $s$ rows of $C$ is a scalar syzygy of $P_2$, we could in principle rewrite $s$ of the polynomials in $P_2$ as linear combinations of the other ${r(d + n - 1)!\over d! (n - 1)!} - s$. Algorithm \ref{alg} takes this fact into account by replacing the $s$ entries of $P_2$ which correspond to the pivot columns of $C$ with zero and calling the result $G$. In constructing $G$, it has isolated the subspace of $S_0(P_2)$ which is in correspondence with the new degree two syzygies of $P_0$ (if any new degree two syzygies exist). Next, Algorithm \ref{alg} uses Subroutine \ref{sub2} to determine a basis, $D$, for the subspace of $S_0(P_2)$ under consideration. However, in order to search for degree three syzygies in an analogous fashion, we need a basis for $S_0(P_2)$ in its entirety. Therefore, Algorithm \ref{alg} sets $B_2 = C \bigcup D$ (for the sake of discussion assuming that Subroutine \ref{sub2} found new degree two syzygies). Finally, Algorithm \ref{alg} solves for the new degree two syzygies themselves by applying Subroutine \ref{sub3} to $D$. After incrementing $d$ to three, Algorithm \ref{alg} would pass through the upper branch of the If statement again in an attempt to find new degree three syzygies of $P_0$.
Of course, we have been assuming throughout our discussion that $\Delta > 3$. It is worth emphasizing that, the termination condition we are using for Algorithm \ref{alg} comes entirely from the physics. The fact that irreducible numerators typically have, at worst, a relatively small mass dimension is the only reason that we were able to fruitfully apply the ideas of reference~\cite{LASyz} and construct Algorithm \ref{alg}. Although we believe that the treatment given here is appropriate, some readers may prefer a more formal one. If that is the case, then we recommend reading~\cite{LASyz}. Most of the non-trivial aspects of our pseudo-code are treated there as well in the style preferred by mathematicians.
\section{An Explicit Example}
\label{example}
In this section, we show how Algorithm \ref{alg} works in practice by going through it for a particular example. It was challenging to find a physically motivated example compact enough to present in detail and, at the same time, rich enough to give the reader a good feeling for how the algorithm functions. In the end, we found that the module given by the irreducible part of the module associated (associated in the sense of the construction reviewed in Section \ref{review}) to the planar massless double box works very well. In the solution algorithm of GKK, the study of this module (hereafter referred to as $\mathcal{M}$) is the first step towards the determination of the complete set\footnote{Here it is perhaps worth pointing out that, in fact, the application of our algorithm to the module associated to the planar massless double box yields more independent solutions than GKK found (working modulo reducibility as they do). We conjecture that, perhaps, GKK did not really seek the {\it complete} set of linearly independent syzygies modulo reducibility but were instead content to determine a subset sufficient for the elimination of as many irreducible numerators as possible. If our reading of GKK is correct, then the obvious question is whether discarding potentially useful linear relations between Feynman integrals of a given topology is prudent. We suspect that this is not the best strategy because, at least for the planar double box, the elimination of irreducible numerators in this manner seems to lead to a large number of irreducible integrals of simpler topology. Unfortunately, further exploration of this interesting question is beyond the scope of the present paper.} of linearly independent syzygies of the module associated to the planar massless double box. In their paper, GKK assert that $\mathcal{M}$ has just three linearly independent syzygies: one of degree one and two of degree two. This {\it a priori} knowledge of the syzygy module of $\mathcal{M}$ will allow us to stop the example when it ceases to be interesting. Otherwise we would have to make several more trips through the While loop (as explained in~\cite{GKK}, $\Delta = 6$ for the planar massless double box), each time discovering no new syzygies. For the sake of clarity, our exposition will mirror the pseudo-code of Section \ref{body} quite closely.
As a preliminary step we must derive the generators of $\mathcal{M}$. In Section \ref{review}, we pointed out that the ordering adopted by GKK for the propagator denominators of the planar massless double box is given by $\{\ell_1^2, \ell_1^2 - 2\ell_1\cdot k_1, \ell_1^2 - 2\ell_1\cdot k_1 -2 \ell_1\cdot k_2 +s_{1 2}, \ell_2^2, \ell_2^2 - 2 \ell_2\cdot k_4, \ell_2^2 - 2 \ell_2\cdot k_3 - 2 \ell_2\cdot k_4 + s_{12},\ell_1^2 + \ell_2^2 + 2 \ell_1\cdot \ell_2\}$. We actually prefer the ordering $\{\ell_1^2, \ell_1^2 - 2\ell_1\cdot k_1, \ell_1^2 - 2\ell_1\cdot k_1 -2 \ell_1\cdot k_2 +s_{1 2}, \ell_2^2 - 2 \ell_2\cdot k_3 - 2 \ell_2\cdot k_4 + s_{12}, \ell_2^2 - 2 \ell_2\cdot k_4, \ell_2^2,\ell_1^2 + \ell_2^2 + 2 \ell_1\cdot \ell_2\}$ and this is what we will use. Note, however, that we do adopt their ordering for the generators themselves. By definition, the generators of $\mathcal{M}$ are the generators of the module associated to the planar massless double box by eqs. (\ref{IBP4}) (the rows of the matrix $E$ in eq. (5.3) of~\cite{GKK}) reduced over the propagator denominators of the massless double box. This reduction is effected by making the substitutions $\{k_3 \rightarrow - k_1 - k_2 - k_4, \ell_1^2 \rightarrow 0, \ell_1 \cdot k_1 \rightarrow 0, \ell_2^2 \rightarrow 0, \ell_2 \cdot k_4 \rightarrow 0, \ell_1 \cdot \ell_2 \rightarrow 0, \ell_1\cdot k_2 \rightarrow s_{12}/2,\ell_2\cdot k_2 \rightarrow -\ell_2\cdot k_1 - s_{12}/2\}$. We find that $\mathcal{M}$ is generated by
\bea
&&\bigg\{\left(0,0,-\frac{s_{12}}{2},0,0,0,0\right), \left(0,-\ell_2\cdot k_1,\frac{s_{12}}{2},0,0,0,0\right), \left(0,0,-\frac{s_{12}}{2},0,0,0,\ell_2\cdot k_1\right),
\nonumber \\ &&
\left(\frac{s_{12}}{2},0,0,0,0,0,-\ell_2\cdot k_1\right), \left(\ell_1\cdot k_4,-\frac{\chi_{14}s_{12}}{2}+\ell_1\cdot k_4,\frac{s_{12}}{2}+\ell_1\cdot k_4,0,0,0,\ell_1\cdot k_4\right),
\nonumber \\ &&
\left(0,0,0,\frac{s_{12}}{2},-\ell_1\cdot k_4,0,0\right), \left(0,0,0,-\frac{s_{12}}{2},0,0,0\right),\left(0,0,0,\frac{s_{12}}{2}+\ell_2\cdot k_1,-\frac{\chi_{14}s_{12}}{2}\right.
\nonumber \\ &&
\left. + \ell_2\cdot k_1,\ell_2 \cdot k_1, \ell_2\cdot k_1\right),\left(0,0,0,-\ell_2\cdot k_1,\frac{\chi_{14}s_{12}}{2}-\ell_2\cdot k_1,-\frac{s_{12}}{2}-\ell_2\cdot k_1,-\ell_2\cdot k_1\right),
\nonumber \\ &&
\left(0,0,0,-\frac{s_{12}}{2},0,0,\ell_1\cdot k_4\right)\bigg\}\,.
\label{gen1}
\eea
If we let $x_1 = \ell_1\cdot k_4$, $x_2 = \ell_2\cdot k_1$, $x_3 = s_{12}$, and take the dot product of each generator in (\ref{gen1}) with $\left(t_1,t_2,t_3,t_4,t_5,t_6,t_7\right)$, we arrive at the generating set for the ideal (hereafter referred to as $\mathcal{I}$) canonically associated to $\mathcal{M}$, $P_0$:
\bea
&&P_0 = \left\{-\frac{1}{2} t_3 x_3, \frac{t_3 x_3}{2}-t_2 x_2,t_7 x_2-\frac{t_3 x_3}{2}, \frac{t_1 x_3}{2}-t_7 x_2, -\frac{1}{2} t_2 x_3 \chi _{14}+t_1 x_1+t_2 x_1+t_3 x_1\right.
\nonumber \\ &&
\left.+t_7 x_1+\frac{t_3 x_3}{2}, \frac{t_4 x_3}{2}-t_5 x_1, -\frac{1}{2} t_4 x_3,-\frac{1}{2} t_5 x_3 \chi _{14}+t_4 x_2+t_5 x_2+t_6 x_2+t_7 x_2+\frac{t_4
x_3}{2},\right.
\nonumber \\ &&
\left.\frac{1}{2} t_5 x_3 \chi _{14}-t_4 x_2-t_5 x_2-t_6 x_2-t_7 x_2-\frac{t_6 x_3}{2},t_7 x_1-\frac{t_4 x_3}{2}\right\}\,.
\label{polys}
\eea
Only the first ten rows of $E$ remain non-zero after the reduction is carried out. Provided that all $m$ propagator denominators are independent of one another (which is certainly true in our case), the $P_0$ we arrive at in this fashion will always be a minimal generating set for $\mathcal{I}$ which implies that, as assumed in Algorithm \ref{alg}, ${\rm SyzList}_0 = B_0 = \{{\bf 0}\}$. Actually, for most of steps of the algorithm the explicit form of $P_0$ is unimportant and we suppress it, writing instead $P_0 = \{p_1,p_2,p_3,p_4,p_5,p_6,p_7,p_8,p_9,p_{10}\}$.
We initialize $d$ to one and enter the While loop. Since $B_0 = \{{\bf 0}\}$, the algorithm directs us to the lower branch of the If statement. $M_1^{(3)} = \{x_1,x_2,x_3\}$, $P_0 = \{p_1,p_2,p_3,p_4,p_5,p_6,p_7,$
$p_8,p_9,p_{10}\}$, and
\begin{changemargin}{-.7 cm}{0 cm}
\bea
&&P_1 = \left\{p_1 x_1,p_1 x_2,p_1 x_3,p_2 x_1, p_2 x_2, p_2 x_3, p_3 x_1, p_3 x_2, p_3 x_3, p_4 x_1, p_4 x_2, p_4 x_3, p_5 x_1, p_5 x_2, p_5 x_3,\right.
\nonumber \\ &&
\left.p_6 x_1,p_6 x_2,p_6 x_3, p_7 x_1, p_7 x_2, p_7 x_3, p_8 x_1, p_8 x_2, p_8 x_3, p_9 x_1, p_9 x_2, p_9 x_3, p_{10} x_1, p_{10} x_2, p_{10} x_3\right\}\nonumber \\
\eea
\end{changemargin}
has ${10 (1+3-1)!\over 1!(3-1)!}= 30$ elements. To find $D$ (Subroutine \ref{sub2}) we have to introduce a sequence of 30 dummy variables
\bea
Y &=& \left\{y_1, y_2, y_3, y_4, y_5, y_6, y_7, y_8, y_9, y_{10}, y_{11}, y_{12}, y_{13}, y_{14}, y_{15}, y_{16},\right.
\nonumber \\ &&
\left.y_{17}, y_{18}, y_{19}, y_{20}, y_{21}, y_{22}, y_{23}, y_{24}, y_{25}, y_{26}, y_{27}, y_{28}, y_{29}, y_{30}\right\}\,,
\eea
expand the dot product $G \cdot Y$, and partition the terms in the resulting sum into equivalence classes. Two terms are equivalent if and only if they contain the same monomial in the variables $\{t_1,t_2,t_3,t_4,t_5,t_6,t_7,x_1,x_2,x_3\}$. Using the explicit form of the $p_i$ given in eq. (\ref{polys}), we find 37 equivalence classes:
\begin{changemargin}{-.8 cm}{0 cm}
\bea
&&\left\{t_1 x_1^2 y_{13},t_2 x_1^2 y_{13},t_3 x_1^2 y_{13},-t_5 x_1^2 y_{16},t_7 y_{13} x_1^2+t_7 y_{28} x_1^2,t_1 x_1 x_2 y_{14},t_2 x_1 x_2 y_{14}-t_2 x_1 x_2 y_4,\right.
\nonumber \\ &&
\left.t_3 x_1 x_2 y_{14},t_4 x_1 x_2 y_{22}-t_4 x_1 x_2 y_{25},-t_5 x_1 x_2 y_{17}+t_5 x_1 x_2 y_{22}-t_5 x_1 x_2 y_{25},t_6 x_1 x_2 y_{22}-t_6 x_1 x_2 y_{25},\right.
\nonumber \\ &&
\left. t_7 x_1 x_2 y_7-t_7 x_1 x_2 y_{10}+t_7 x_1 x_2 y_{14}+t_7 x_1 x_2 y_{22}-t_7 x_1 x_2 y_{25}+t_7 x_1 x_2 y_{29},-t_2 x_2^2 y_5,t_4 x_2^2 y_{23}\right.
\nonumber \\ &&
\left.-t_4 x_2^2 y_{26},t_5 x_2^2 y_{23}-t_5 x_2^2 y_{26},t_6 x_2^2 y_{23}-t_6 x_2^2 y_{26},t_7 y_8 x_2^2-t_7 y_{11} x_2^2+t_7 y_{23} x_2^2-t_7 y_{26} x_2^2,\right.
\nonumber \\ &&
\left.\frac{1}{2} t_1 x_1 x_3
y_{10}+t_1 x_1 x_3 y_{15},t_2 x_1 x_3 y_{15}-\frac{1}{2} t_2 x_1 x_3 y_{13} \chi _{14},-\frac{1}{2} t_3 x_1 x_3 y_1+\frac{1}{2} t_3 x_1 x_3
y_4\right.
\nonumber \\ &&
\left.-\frac{1}{2} t_3 x_1 x_3 y_7+\frac{1}{2} t_3 x_1 x_3 y_{13}+t_3 x_1 x_3 y_{15},\frac{1}{2} t_4 x_1 x_3 y_{16}-\frac{1}{2} t_4 x_1 x_3
y_{19}+\frac{1}{2} t_4 x_1 x_3 y_{22}\right.
\nonumber \\ &&
\left.-\frac{1}{2} t_4 x_1 x_3 y_{28},-t_5 x_1 x_3 y_{18}-\frac{1}{2} t_5 x_1 x_3 y_{22} \chi _{14}+\frac{1}{2} t_5
x_1 x_3 y_{25} \chi _{14},-\frac{1}{2} t_6 x_1 x_3 y_{25},t_7 x_1 x_3 y_{15}\right.
\nonumber \\ &&
\left.+t_7 x_1 x_3 y_{30},\frac{1}{2} t_1 x_2 x_3 y_{11},-t_2 x_2 x_3
y_6-\frac{1}{2} t_2 x_2 x_3 y_{14} \chi _{14},-\frac{1}{2} t_3 x_2 x_3 y_2+\frac{1}{2} t_3 x_2 x_3 y_5\right.
\nonumber \\ &&
\left.-\frac{1}{2} t_3 x_2 x_3 y_8+\frac{1}{2} t_3
x_2 x_3 y_{14},\frac{1}{2} t_4 x_2 x_3 y_{17}-\frac{1}{2} t_4 x_2 x_3 y_{20}+\frac{1}{2} t_4 x_2 x_3 y_{23}+t_4 x_2 x_3 y_{24}\right.
\nonumber \\ &&
\left.-t_4 x_2 x_3
y_{27}-\frac{1}{2} t_4 x_2 x_3 y_{29},t_5 x_2 x_3 y_{24}-t_5 x_2 x_3 y_{27}-\frac{1}{2} t_5 x_2 x_3 y_{23} \chi _{14}+\frac{1}{2} t_5 x_2 x_3
y_{26} \chi _{14},\right.
\nonumber \\ &&
\left.t_6 x_2 x_3 y_{24}-\frac{1}{2} t_6 x_2 x_3 y_{26}-t_6 x_2 x_3 y_{27},t_7 x_2 x_3 y_9-t_7 x_2 x_3 y_{12}+t_7 x_2 x_3
y_{24}-t_7 x_2 x_3 y_{27},\right.
\nonumber \\ &&
\left.\frac{1}{2} t_1 x_3^2 y_{12},-\frac{1}{2} t_2 x_3^2 y_{15} \chi _{14},-\frac{1}{2} t_3 y_3 x_3^2+\frac{1}{2} t_3 y_6 x_3^2-\frac{1}{2} t_3 y_9 x_3^2+\frac{1}{2} t_3 y_{15} x_3^2,\frac{1}{2} t_4 y_{18} x_3^2 \right.
\nonumber \\ &&
\left.-\frac{1}{2} t_4 y_{21} x_3^2+\frac{1}{2} t_4 y_{24}
x_3^2-\frac{1}{2} t_4 y_{30} x_3^2,\frac{1}{2} t_5 x_3^2 y_{27} \chi _{14}-\frac{1}{2} t_5 x_3^2 y_{24} \chi _{14},-\frac{1}{2} t_6 x_3^2 y_{27}\right\}\,.
\eea
\end{changemargin}
After setting the terms in each equivalence class to zero, we find that six of the resulting equations are obviously redundant. The remaining 31 equations,
\bea
&&\left\{0=y_5,0=y_{11},0=y_{12},0=y_{13},0=y_4-y_{14},0=y_2-y_5+y_8-y_{14},0=y_{14},\right.
\nonumber \\ &&
\left.0=\frac{y_{14} \chi_{14}}{2}+y_6,0=y_1-y_4+y_7-y_{13}-2
y_{15},0=y_3-y_6+y_9-y_{15},0=y_{13}-\frac{2 y_{15}}{\chi _{14}},\right.
\nonumber \\ &&
\left.0=y_{15},0=y_{10}+2 y_{15},0=y_{16},0=\frac{y_{22} \chi _{14}}{2}-\frac{y_{25} \chi_{14}}{2}+y_{18},0=y_{22}-y_{25},0=y_{25},\right.
\nonumber \\ &&
\left.0=y_{17}-y_{22}+y_{25},0=y_{23}-y_{26},0=y_8-y_{11}+y_{23}-y_{26},0=y_{24}-y_{27},\right.
\nonumber \\ &&
\left.0=y_9-y_{12}+y_{24}-y_{27},0=y_{24}-\frac{y_{26}}{2}-y_{27},0=-\frac{2 y_{24}}{\chi _{14}}+\frac{2 y_{27}}{\chi_{14}}+y_{23}-y_{26},0=y_{27},\right.
\nonumber \\ &&
\left.0=y_{16}-y_{19}+y_{22}-y_{28},0=y_{13}+y_{28},0=y_{17}-y_{20}+y_{23}+2 y_{24}-2
y_{27}-y_{29},\right.
\nonumber \\ &&
\left.0=y_7-y_{10}+y_{14}+y_{22}-y_{25}+y_{29},0=y_{18}-y_{21}+y_{24}-y_{30},0=y_{15}+y_{30}\right\}\,,
\label{eqs1}
\eea
constrain the $y_i$. After putting (\ref{eqs1}) into matrix form, we can easily solve for the null space of the resulting matrix. We find that, as expected, the null space is one-dimensional:
\be
D = \Big\{\left(1,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,1,0\right)\Big\}.
\ee
We have now determined a basis for the scalar syzygies of $P_1$:
\be
B_1 = D
\ee
To determine what syzygy of $P_0$ this element of $S_0(P_1)$ corresponds to, we apply the same mapping that we used in Section \ref{review} (formalized in Subroutine \ref{sub3}) in the example illustrating why computing a basis for a syzygy module with linear algebra is non-trivial. We partition the single element of $D$ into ten distinct non-overlapping sequences of length three without disturbing the ordering of the vector that we started with
\be
\Big\{\{1,0,0\},\{0,0,0\},\{-1,0,0\},\{0,0,0\},\{0,0,0\},\{0,0,0\},\{0,-1,0\},\{0,0,0\},\{0,0,0\},\{0,1,0\}\Big\}
\ee
and then take the dot product of each sequence of three elements with $M_1^{(3)} = \{x_1,x_2,x_3\}$:
\be
{\rm SyzList}_1 = \bigg\{\Big(x_1,0,-x_1,0,0,0,-x_2,0,0,x_2\Big)\bigg\}\,.
\ee
It is easy to check that $(x_1,0,-x_1,0,0,0,-x_2,0,0,x_2)$ is indeed a syzygy of $P_0$.
Now we increment $d$ to two and begin our second pass through the While loop. This time, $B_1 \neq \{{\bf 0}\}$ and, therefore, the algorithm directs us to the upper branch of the If statement. $M_1^{(3)} = \{x_1,x_2,x_3\}$, $P_0 = \{p_1,p_2,p_3,p_4,p_5,p_6,p_7,p_8,p_9,p_{10}\}$, and
\begin{changemargin}{-.7 cm}{0 cm}
\bea
&&P_1 = \left\{p_1 x_1,p_1 x_2,p_1 x_3,p_2 x_1, p_2 x_2, p_2 x_3, p_3 x_1, p_3 x_2, p_3 x_3, p_4 x_1, p_4 x_2, p_4 x_3, p_5 x_1, p_5 x_2, p_5 x_3,\right.
\nonumber \\ &&
\left.p_6 x_1,p_6 x_2,p_6 x_3, p_7 x_1, p_7 x_2, p_7 x_3, p_8 x_1, p_8 x_2, p_8 x_3, p_9 x_1, p_9 x_2, p_9 x_3, p_{10} x_1, p_{10} x_2, p_{10} x_3\right\}\nonumber \\
\eea
\end{changemargin}
with ${10 (1+3-1)!\over 1!(3-1)!}= 30$ elements. To determine $P_2$, we need the set of monomials of degree two built out of $\{x_1,x_2,x_3\}$, $M_2^{(3)} = \left\{x_1^2,x_1 x_2, x_1 x_3, x_2^2, x_2 x_3, x_3^2\right\}$. As mentioned in Section \ref{body}, for definiteness, we have chosen the lexicographical ordering for our sets of monomials. Let us say again that this choice is not by any means necessary. A different choice of ordering for the monomials would lead to a different representation of the same syzygy module; the number of linearly independent syzygies at each degree would be exactly the same. Continuing, we see that
\bea
&&P_2 = \left\{p_1 x_1^2,p_1 x_1 x_2,p_1 x_1 x_3,p_1 x_2^2,p_1 x_2 x_3,p_1 x_3^2,p_2 x_1^2,p_2 x_1 x_2,p_2 x_1 x_3,p_2 x_2^2,p_2 x_2 x_3,p_2 x_3^2,\right.
\nonumber \\ &&
\left.p_3 x_1^2,p_3 x_1 x_2,p_3 x_1 x_3,p_3 x_2^2,p_3 x_2 x_3,p_3 x_3^2,p_4 x_1^2,p_4 x_1 x_2,p_4 x_1 x_3,p_4 x_2^2,p_4 x_2 x_3,p_4 x_3^2,\right.
\nonumber \\ &&
\left.p_5 x_1^2,p_5 x_1 x_2,p_5 x_1 x_3,p_5 x_2^2,p_5 x_2 x_3,p_5 x_3^2,p_6 x_1^2,p_6 x_1 x_2,p_6 x_1 x_3,p_6 x_2^2,p_6 x_2 x_3,p_6 x_3^2,\right.
\\
&&\left.p_7 x_1^2,p_7 x_1 x_2,p_7 x_1 x_3,p_7 x_2^2,p_7 x_2 x_3,p_7 x_3^2,p_8 x_1^2,p_8 x_1 x_2,p_8 x_1 x_3,p_8 x_2^2,p_8 x_2 x_3,p_8 x_3^2,\right.
\nonumber \\ &&
\left.p_9 x_1^2,p_9 x_1 x_2,p_9 x_1 x_3,p_9 x_2^2,p_9 x_2 x_3,p_9 x_3^2,p_{10} x_1^2,p_{10} x_1 x_2,p_{10} x_1 x_3,p_{10} x_2^2,p_{10} x_2 x_3,p_{10} x_3^2\right\}\nonumber
\eea
has ${10 (2+3-1)!\over 2!(3-1)!}= 60$ elements. $P_1$ is the basis with respect to which the single element of $B_1$ has components
\be
\left(1,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,1,0\right)\,.
\label{deg1syz}
\ee
The function $\sigma$ (Subroutine \ref{sub1}) maps this vector to three new vectors in the larger vector space for which $P_2$
is the standard basis. This map is carried out by extending the $P_1$ by each of the three variables in turn. Applying $\sigma$ to
\bea
&&\left(1,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,1,0\right)
\nonumber \\ &&
\qquad\qquad\qquad= p_1 x_1-p_3 x_1-p_7 x_2+p_{10} x_2\,,
\eea
we obtain
\bea
&&\sigma\left(p_1 x_1-p_3 x_1-p_7 x_2+p_{10} x_2\right) = \left\{p_1 x_1^2-p_3 x_1^2-p_7 x_1 x_2+p_{10}x_1 x_2,\right.
\nonumber \\ &&
\left.p_1 x_1 x_2-p_3 x_1 x_2-p_7 x_2^2+p_{10}x_2^2, p_1 x_1 x_3-p_3 x_1 x_3-p_7 x_2 x_3+p_{10}x_2 x_3\right\}\,.
\eea
We can now read off the components of three scalar syzygies of $P_2$ which have their origin in the scalar syzygy of $P_1$ that we computed on our first pass through the While loop. They are:
\begin{changemargin}{-.7 cm}{0 cm}
\bea
&&A =\Big\{(1,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0,0,0,0), (0,1,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,
\nonumber \\ &&
0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,1,0,0), (0,0,1,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0)\Big\}\,.
\eea
\end{changemargin}
If we view the elements of $A$ as the rows of a matrix, the result is already in row echelon form. It follows that $C$ is simply $A$ in matrix form. The first three columns of $C$ are pivot columns, so $G$ is $P_2$ with the first three entries replaced with zero:
\begin{changemargin}{-.6 cm}{0 cm}
\bea
&&G = \left\{0,0,0,p_1 x_2^2,p_1 x_2 x_3,p_1 x_3^2,p_2 x_1^2,p_2 x_1 x_2,p_2 x_1 x_3,p_2 x_2^2,p_2 x_2 x_3,p_2 x_3^2,p_3 x_1^2,\right.
\nonumber \\ &&
\left.p_3 x_1 x_2,p_3 x_1 x_3,p_3 x_2^2,p_3 x_2 x_3,p_3 x_3^2,p_4 x_1^2,p_4 x_1 x_2,p_4 x_1 x_3,p_4 x_2^2,p_4 x_2 x_3,p_4 x_3^2,p_5 x_1^2,\right.
\nonumber \\ &&
\left.p_5 x_1 x_2,p_5 x_1 x_3,p_5 x_2^2,p_5 x_2 x_3,p_5 x_3^2,p_6 x_1^2,p_6 x_1 x_2,p_6 x_1 x_3,p_6 x_2^2,p_6 x_2 x_3,p_6 x_3^2,p_7 x_1^2,\right.
\\
&&\left.p_7 x_1 x_2,p_7 x_1 x_3,p_7 x_2^2,p_7 x_2 x_3,p_7 x_3^2,p_8 x_1^2,p_8 x_1 x_2,p_8 x_1 x_3,p_8 x_2^2,p_8 x_2 x_3,p_8 x_3^2,p_9 x_1^2,\right.
\nonumber \\ &&
\left.p_9 x_1 x_2,p_9 x_1 x_3,p_9 x_2^2,p_9 x_2 x_3,p_9 x_3^2,p_{10} x_1^2,p_{10} x_1 x_2,p_{10} x_1 x_3,p_{10} x_2^2,p_{10} x_2 x_3,p_{10} x_3^2\right\}\nonumber\,.
\eea
\end{changemargin}
Making this replacement will prevent Subroutine \ref{sub2} from rediscovering syzygies that have their origin in the syzygy of $P_1$ that we found on our first pass through the While loop. On this pass through the loop, we refrain from stepping into Subroutines \ref{sub2} and \ref{sub3} since their functionality should already be quite clear from our first pass through. Applying ${\rm SSyz}$ to $G$, we find
\begin{changemargin}{-.5 cm}{0 cm}
\bea
&&D = {\rm SSyz}(G) = \bigg\{\Big(0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,-1,\frac{\chi_{14}}{2},0,-4,-1,0,-2,\frac{\chi_{14}}{2},0,-2,-1,0,0,0,0,-2,0,0,0,0,0,0,0,0,1,0\Big), \Big(0,0,0,0,
\nonumber \\ &&
-\frac{1}{4},\frac{\chi_{14}}{8},0,0,-\frac{1}{4},0,0,\frac{\chi_{14}}{8},0,-\frac{1}{2},-\frac{1}{4},0,0,0,0,\frac{1}{2},0,0,0,0,0,0,0,0,-\frac{1}{4},0,0,0,0,0,\frac{1}{2},-\frac{\chi_{14}}{4},
\nonumber \\ &&
0,2,\frac{1}{2},-1,\frac{1}{2},-\frac{\chi _{14}}{4},0,1,\frac{1}{2},0,0,0,0,1,0,0,0,0,0,0,0,1,0,0\Big)\bigg\}
\eea
\end{changemargin}
for the new scalar syzygies of $P_2$. Since $D \neq \{{\bf 0}\}$,
\begin{changemargin}{-.5 cm}{0 cm}
\bea
&&B_2 = C \cup D = \Big\{(1,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0,0,0,0), (0,1,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,0,0,1,0,0), (0,0,1,0,0,0,0,0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,0,0,0,0,0,-1,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0), \Big(0,0,0,0,0,0,0,0,0,0,0,0,0,
\nonumber \\ &&
0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,-1,\frac{\chi_{14}}{2},0,-4,-1,0,-2,\frac{\chi_{14}}{2},0,-2,-1,0,0,0,
\nonumber \\ &&
0,-2,0,0,0,0,0,0,0,0,1,0\Big),\Big(0,0,0,0,-\frac{1}{4},\frac{\chi_{14}}{8},0,0,-\frac{1}{4},0,0,\frac{\chi_{14}}{8},0,-\frac{1}{2},-\frac{1}{4},0,0,0,0,\frac{1}{2},0,
\nonumber \\ &&
0,0,0,0,0,0,0,-\frac{1}{4},0,0,0,0,0,\frac{1}{2},-\frac{\chi_{14}}{4},0,2,\frac{1}{2},-1,\frac{1}{2},-\frac{\chi_{14}}{4},0,1,\frac{1}{2},0,0,0,0,1,0,0,0,0,0,0,
\nonumber \\ &&
0,1,0,0\Big)\Big\}\,.
\eea
\end{changemargin}
The two new scalar syzygies of $P_2$ discovered by {\rm SSyz} map, via {\rm ToSyz}, to two new syzygies of $P_0$ of degree two, linearly independent of $(x_1,0,-x_1,0,0,0,-x_2,0,0,x_2)$:
\bea
&&{\rm SyzList}_2 = {\rm ToSyz}(D) = \bigg\{\Big(0,0,0,0,0,\frac{1}{2} x_3^2 \chi _{14}-x_2 x_3,\frac{1}{2} x_3^2 \chi _{14}-x_1 x_3-2 x_2 x_3
\nonumber \\ &&
-4 x_1 x_2,-2 x_1 x_2-x_1 x_3,-2 x_1 x_2,x_2 x_3\Big),\Big(\frac{1}{8} x_3^2 \chi _{14}-\frac{x_2 x_3}{4},\frac{1}{8} x_3^2 \chi _{14}-\frac{x_1 x_3}{4},
\nonumber \\ &&
-\frac{1}{2} x_1 x_2-\frac{x_1 x_3}{4},\frac{x_1 x_2}{2},-\frac{1}{4} x_2 x_3,\frac{x_2 x_3}{2}-\frac{1}{4} x_3^2 \chi _{14},-\frac{1}{4} x_3^2 \chi_{14}-x_2^2+2 x_1 x_2
\nonumber \\ &&
+\frac{x_3 x_2}{2}+\frac{x_1 x_3}{2},x_1 x_2+\frac{x_1 x_3}{2},x_1 x_2,x_2^2\Big)\bigg\}\,.
\eea
It is again simple to check that the elements of ${\rm SyzList}_2$ are in fact syzygies of $P_0$. If we did not have {\it a priori} knowledge of the syzygy module, we would have to continue on and pass through the While loop several more times (as explained in~\cite{GKK}, $\Delta = 6$ for the planar massless double box) to search for more potentially useful linearly independent syzygies.\footnote{It is important to realize that, in general, $D = \{{\bf 0}\}$ for $d = i$ does {\it not} imply that $D = \{{\bf 0}\}$ for $d = i + 1$, though this may turn out to be true in practice for physically motivated examples.} Since we do not expect to find further solutions, we can exit the While loop after just two iterations and collect the results:
\bea
&&\bigcup\limits_{d=1}^2 {\rm SyzList}_d = \bigg\{\Big(x_1,0,-x_1,0,0,0,-x_2,0,0,x_2\Big),\Big(0,0,0,0,0,\frac{1}{2} x_3^2 \chi _{14}-x_2 x_3,\frac{1}{2} x_3^2 \chi_{14}
\nonumber \\ &&
-x_1 x_3-2 x_2 x_3-4 x_1 x_2,-2 x_1 x_2-x_1 x_3,-2 x_1 x_2,x_2 x_3\Big),\Big(\frac{1}{8} x_3^2 \chi _{14}-\frac{x_2 x_3}{4},\frac{1}{8} x_3^2 \chi _{14}
\nonumber \\ &&
-\frac{x_1 x_3}{4},-\frac{1}{2} x_1 x_2-\frac{x_1 x_3}{4},\frac{x_1 x_2}{2},-\frac{1}{4} x_2 x_3,\frac{x_2 x_3}{2}-\frac{1}{4} x_3^2 \chi _{14},-\frac{1}{4} x_3^2 \chi_{14}-x_2^2+2 x_1 x_2
\nonumber \\ &&
+\frac{x_3x_2}{2}+\frac{x_1 x_3}{2},x_1 x_2+\frac{x_1 x_3}{2},x_1 x_2,x_2^2\Big)\bigg\}\,.
\eea
Before leaving this section, we translate the above result back into the usual language used to describe the planar massless double box:
\begin{changemargin}{-.7 cm}{0 cm}
\bea
&&\bigcup\limits_{d=1}^2 {\rm SyzList}_d = \bigg\{\Big(\ell_1 \cdot k_4,0,-\ell_1 \cdot k_4,0,0,0,-\ell_2 \cdot k_1,0,0,\ell_2 \cdot k_1\Big),\Big(0,0,0,0,0,\frac{1}{2} s_{12}^2 \chi _{14}
\\
&&-\ell_2\cdot k_1\,s_{12},\frac{1}{2} s_{12}^2 \chi_{14}-\ell_1 \cdot k_4 \,s_{12}-2 \ell_2 \cdot k_1\, s_{12}-4 \ell_1 \cdot k_4\, \ell_2 \cdot k_1,-2 \ell_1 \cdot k_4\, \ell_2 \cdot k_1-\ell_1 \cdot k_4\, s_{12},
\nonumber \\ &&
-2 \ell_1 \cdot k_4\, \ell_2 \cdot k_1,\ell_2 \cdot k_1\, s_{12}\Big),\Big(\frac{1}{8} s_{12}^2 \chi _{14}-\frac{\ell_2 \cdot k_1\, s_{12}}{4},\frac{1}{8} s_{12}^2 \chi _{14}-\frac{\ell_1 \cdot k_4\, s_{12}}{4},-\frac{1}{2} \ell_1 \cdot k_4 \,\ell_2 \cdot k_1
\nonumber \\ &&
-\frac{\ell_1 \cdot k_4\, s_{12}}{4},\frac{\ell_1 \cdot k_4 \,\ell_2 \cdot k_1}{2},-\frac{1}{4} \ell_2 \cdot k_1\, s_{12},\frac{\ell_2 \cdot k_1\, s_{12}}{2}-\frac{1}{4} s_{12}^2 \chi _{14},-\frac{1}{4} s_{12}^2 \chi_{14}-\left(\ell_2 \cdot k_1\right)^2
\nonumber \\
&&+2 \ell_1 \cdot k_4\, \ell_2 \cdot k_1+\frac{s_{12}\,\ell_2 \cdot k_1}{2}+\frac{\ell_1 \cdot k_4\, s_{12}}{2},\ell_1 \cdot k_4\, \ell_2 \cdot k_1+\frac{\ell_1 \cdot k_4\, s_{12}}{2},\ell_1 \cdot k_4\,\ell_2 \cdot k_1,\left(\ell_2 \cdot k_1\right)^2\Big)\bigg\}\,. \nonumber
\eea
\end{changemargin}
\section{Conclusions}
\label{ConclusionSection}
This paper continued the program of research initiated in~\cite{GKK} by Gluza, Kajda, and Kosower (GKK). In reference~\cite{GKK}, GKK proposed a novel reduction scheme for multi-loop integrals guaranteed to produce unitarity-compatible integral bases free of doubled propagator denominators. In this work we took a fresh look at the computationally intensive part of their procedure, the generation of complete sets of unitarity-compatible integration by parts relations. Drawing upon some of the ideas in reference~\cite{LASyz}, we found an attractive alternative to the GKK approach which completely avoids the use of Gr\"{o}bner bases. In fact, we showed in Section \ref{body} that our solution, Algorithm \ref{alg}, can be described in terms of simple linear algebra.
One shortcoming of the present paper is that we cannot yet claim to have fully optimized Algorithm \ref{alg}. Even if we excise the trivial redundancies introduced into our pseudo-code for the sake of clarity, there are many features of the algorithm which might benefit from further optimization. For example, as should already be clear from the non-trivial example discussed in Section \ref{example}, the matrices produced by Algorithm \ref{alg} are typically quite sparse and, so far, we have made no attempt to exploit this feature of the problem. Furthermore, it seems likely that a high-level implementation in {\tt Mathematica} (such as the one written by the author) will ultimately prove insufficient for research purposes. Besides the fact that, as a general rule, {\tt Mathematica} runs very slowly, it is not at all clear that {\tt Mathematica} exploits the best available algorithms for row reducing matrices; if our preliminary experimentations are any guide, it appears that {\tt Mathematica} manages system resources rather poorly. Fortunately, we anticipate that the brevity and simplicity of Algorithm \ref{alg} will make it possible to optimize and then effectively implement at a lower level in {\tt C++} or {\tt Fortran}. We are excited by this prospect and hope to pursue a project along these lines in the near future.
\section*{Acknowledgments}
I would like to thank Janusz Gluza and David Kosower for inspiring me to pursue this line of research. I am also very grateful to Valery Yundin for his critical reading of an earlier draft of this work. The {\tt LaTeX} packages {\sc algorithm} and {\sc algorithmic}, written by Peter Williams and maintained by Roger\'{i}o Brito, were used to typeset Algorithm \ref{alg} and Subroutines \ref{sub1}, \ref{sub2}, and \ref{sub3}. Finally, I gratefully acknowledge CICYT support through the project FPA-2009-09017 and CAM support through the project HEPHACOS S2009/ESP-1473.
\bibliographystyle{JHEP}
|
\section{Introduction}
In the text \cite{paper01} the author has given a new framework to define
$(n-1)$-manifolds in $\mathbb{Z}^n$ together with a notion of ``good pairs'' of adjacency relations.
Such a good pair makes it possible for a $(n-1)$-manifold to satisfy a discrete
analog of the Theorem of Jordan-Brouwer. This Theorem is a generalization of the
Jordan-curve Theorem, which states that every simple closed curve in $\mathbb{R}^2$ separates its
complement in exactly two connected components and is itself the boundary of both of them.
Brouwer showed that the statement is true for simple $(n-1)$-manifolds in $\mathbb{R}^n$ for all
$n\ge2$. It has been an open question since the beginnings of digital image analysis, if this
is true in a discrete setting, so to speak in $\mathbb{Z}^n$.
As the figure \ref{Pic:ausgangs_problem} shows, it is not even clear what a simple
closed curve should look like in a discrete setting. And really, this depends on
the adjacency we impose
on the points of $\mathbb{Z}^n$. We also see from the figure, that it is not enough to use only
one adjacency for the base-set (background / white points) and the objects
(foreground / black points), we have to use pairs of them. Unfortunately, not every pair of
adjacencies is suitable because some even fail to make a ($n-1$)-manifold out of the neighbors
of a given point, and so they do not even satisfy the Theorem of Jordan-Brouwer. On these grounds
the notion of a good pair arose and good pairs are the central topic of this article.
A solution for the points in the figure would be, to equip the black points with the 8-adjacency
and use the 4-adjacency for the white ones. Then is clear that a discrete notion of the
Jordan-theorem is true for this example.
\begin{figure}[htb]
\begin{center}
\includegraphics{eps_files/ausgangs_problem.eps}
\end{center}
\caption[Initial problem]{\small Depending on the adjacency relations we use for the
black and the white points, respectively, the set of black points is connected
(8-adjacency) or disconnected (4-adjacency). Also the set of white points may be connected
(8-adjacency) or disconnected (4-adjacency). Only 4-adjacency is depicted.}
\label{Pic:ausgangs_problem}
\end{figure}
For a long time adjacencies like the 4- and 8-neighborhood have been used, and of course, it
is possible to generalize them to higher dimensions. This is done in this paper and we will
see, which pairs of such relations give us good pairs. To do so, we will use the gridcube model
of $\mathbb{Z}^n$ which is widely accepted and may be found in the book of A.~Rosenfeld and R.~Klette
\cite{klette}. It gives us a basic understanding of how these adjacencies may be build in high
dimensions and once we have a good mathematical description for them, we may use it for the study
of pairs of the adjacencies that we will call ``cubical'' because of the relation to this
model.
In the 1980s E.~Khalimsky \cite{khalimsky} proposed a topological motivated approach with
the so called Khalimsky-neighborhood. This topological notion gives also rise to graph-theoretic
adjacencies and so it seems interesting to study it. Since it is already known, that these
relations form good pairs, as seen in \cite{khalimsky} and \cite{herman}, we can use it as a
test for the theory that also shows, how we are able to combine topological and
graph-theoretic concepts.
The paper is organized as follows: We start with some basic definitions in section 2 where
we do a tour through basic discrete topology and the
graph-theoretic knowledge we use in this text, in section 3 the important concepts of
the paper \cite{paper01} are given and in section 4 we apply the theory to the aforementioned
adjacency relations. We end the text with some conclusions in section 5.
\section{Basic Definitions}\label{chapter_4}
\subsection{Topological Basics}
We use this section to introduce some basic topological notions. These stem from the usual
set-theoretic topology as it might be found in any textbook on topology like
the one of St\"ocker and Zieschang \cite{sundz}, but we also introduce some facts
given by P.S.~Alexandrov in his
text \cite{alexandroff}\footnote{Actually, Paul~Alexandroff is the same person as
Pavel~Sergeyevitch~Alexandrov. The different names origin in a different transcription of the
cyrillic letters in German and English.}.
\begin{Def}
A pair $(\mathcal{P},\mathscr{T})$ is called \DF{topological space} for a set $\mathcal{P}$
and set $\mathscr{T}\subset\mathfrak{P}(\mathcal{P})$, the so called \DF{open sets} or \DF{topology} on $\mathcal{P}$,
with the following properties:
\begin{enumerate}
\item $O_i\in\mathscr{T},i\in I \Rightarrow \bigcup_{i\in I}O_i \in \mathscr{T}$
\item $O_1,\ldots,O_n\in\mathscr{T} \Rightarrow \bigcap_{1=1}^n O_i\in\mathscr{T}$
\item $\mathcal{P},\emptyset\in\mathscr{T}$
\end{enumerate}
\end{Def}
A trivial topology on $\mathcal{P}$ is the \DF{discrete} topology
$\mathfrak{P}(\mathcal{P})$. Please do not mistake the special ``discrete'' topology with the
``discrete'' setting we are working in. Even the $\mathbb{R}^n$ may be equipped with a discrete
topology and almost none of the discrete topologies we are referring to in this text are
powersets of the base-set.
The subsets of $\mathcal{P}$, which have an open complement are called \DF{closed}.
An open set $U\in\mathscr{T}$ is called \DF{neighborhood} of a point $x\in\mathcal{P}$ if
$x$ is contained in $U$.
A topological space that satisfies the following stronger claim instead of
property (2), is called \DF{Alexandrov-space}
\begin{enumerate}
\item[2'.] $O_i\in\mathscr{T},i\in I$ $\Rightarrow$ $\bigcap_{i\in I}O_i\in\mathscr{T}.$
\end{enumerate}
All results for topological spaces are also true in Alexandrov-spaces.
Topological spaces may be classified concerning the following separation
properties:
\begin{Def}
A topological space may satisfy some of the separation axioms:
\begin{description}
\item[$T_0$:] $\bigwedge_{x,y\in\mathcal{P}},x\neq y\bigvee_{U\in\mathscr{T}}:(x\in
U\not\ni y)\lor ( x\not\in U\ni y)$
\item[$T_1$:] $\bigwedge_{x,y\in\mathcal{P}},x\neq y\bigvee_{U\in\mathscr{T}}:x\in
U\not\ni y$
\item[$T_2$:] $\bigwedge_{x,y\in\mathcal{P}},x\neq y\bigvee_{U,V\in\mathscr{T}}:(x\in
U\not\ni y)\land ( x\not\in V\ni y)\land (U\cap V=\emptyset)$
\end{description}
\end{Def}
One can see, that every $T_i$-space is also a $T_{i-1}$-space. It is also true,
that considering property (2') interesing only for $T_0$-spaces:
\begin{Lemma}
An Alexandrov-space that satisfies the separation axioms $T_1$ or $T_2$ necessarily
has the discrete topology.
\end{Lemma}
\begin{Beweis}
Let $\mathcal{P}$ be a $T_1$-space, $p\in \mathcal{P}$ and $U$ a neighborhood of $p$. If
$U=\{p\}$, then we are done. Otherwise, there exists a $q\neq p$ in $U$
and by property $(1)$, we may find a neighborhood $U'$, the contains $p$ but not $q$.
The intersection of all these sets is open and so, $\mathcal{P}$ has to be discrete.
The proof for $T_2$-spaces is analog.
\end{Beweis}
To give a topology on a set $\mathcal{P}$, it is enough to give a certain family $\mathcal{B}$
of open sets that can be used to generate all the open sets of $\mathcal{P}$ by using
set-theoretic union.
This family is then called \DF{base} of the topology $\mathscr{T}$.
A topological space is called \DF{locally finite}, if for any point $p$ in $\mathcal{P}$ exists a
finite open set and a finite closed set that both contain $p$.
In the following, we define how we can build new topological spaces from given ones.
\begin{Def}
Let $(P_i,\mathscr{T}_i)$, $i\in I$, be a family of topological spaces and let
$\mathcal{P}=\prod_{i\in I} \mathcal{P}_i$ be their product and $p_i:\mathcal{P}\to \mathcal{P}_i$ projections. The
\DF{product topology} $\mathscr{T}$ is defined by the base
\[\mathcal{B}=\left\{\bigcap_{k\in K}p_k^{-1}(O_k):
O_k\in\mathscr{T}_k,K\subset I, K\mbox{ finite}\right\}\enspace.\]
The space $(\mathcal{P},\mathscr{T})$ is called \emph{topological product} of the
$(\mathcal{P}_i,\mathscr{T}_i)$.
\end{Def}
\begin{Def}
Let $(\mathcal{P},\mathscr{T})$ be a topological space and $A\subset\mathcal{P}$. With the topology
\[\mathscr{T}|_A=\ASET{O\cap A: O\in\mathscr{T}}\]
The set $A$ can be turned into a topological space $(A,\mathscr{T}|_A)$. The topology $\mathscr{T}|_A$
is called \DF{subspace topology} of $A$ with respect $\mathcal{P}$.
\end{Def}
\begin{Def}
A mapping $f:\mathcal{P}\to\mathcal{Q}$ between two topological spaces
$(\mathcal{P},\mathscr{T})$, $(\mathcal{Q},\mathscr{U})$ is called \DF{continuous}, if
for every $O\in\mathscr{U}$ the set $f^{-1}(O)$ is in $\mathscr{T}$.
\end{Def}
\begin{Def}
A topological space $(X,\mathscr{T})$ is called \DF{connected}, if it cannot be
decomposed into two nonempty open sets:
\[\mathcal{P}=O_1\cup O_2,\, O_1,O_2\in\mathscr{T},\, O_1\neq\emptyset\neq O_2 \Rightarrow
O_1\cap O_2\neq\emptyset.\]
A set $A\subset \mathcal{P}$ is called connected, if it is connected in the subspace topology.
\end{Def}
\begin{Lemma}\label{stetigerzshg}
Let $(\mathcal{P},\mathscr{T})$ be a connected topological space and let $(\mathcal{Q},\mathscr{T}')$ be a topological
space. If $f:\mathcal{P}\to\mathcal{Q}$ is a continuous mapping, then $\mathcal{Q}$
is connected.\hspace{\stretch{1}}$\Box$
\end{Lemma}
From the continuity of the projections $p_i$ in the definition of the product topology we can
deduce the following
\begin{Lemma}\label{prodzshg}
A topological space is connected if and only if all of its factors are connected. \hspace{\stretch{1}}$\Box$
\end{Lemma}
We define a \DF{path} of length $m\in\mathbb{N}$ to be a continuous mapping
$w:\SET{0}{m}\to \mathcal{P}$. A path is \DF{closed} if $w(0)=w(m)$.
\begin{Def}
A topological space $(\mathcal{P},\mathscr{T})$ is called \DF{path-connected}, if for any two points
$p,q\in \mathcal{P}$ exists a path $w$ of length $m$ depending only on $p$ and $q$, such
that $w(0)=p$ and $w(m)=q$.
The topological space $(\mathcal{P},\mathscr{T})$ is called \DF{locally path-connected}
if for every point $p\in\mathcal{P}$ and every neighborhood $U$ of $p$ a path-connected
neighborhood $V\subset U$ exists.
\end{Def}
\begin{Folg} The following holds:
\begin{enumerate}
\item Path-connected spaces are connected.
\item Connected and locally path-connected spaces are path-connected.\hspace{\stretch{1}}$\Box$
\end{enumerate}
\end{Folg}
\begin{Def}
Let $X$ and $Y$ be topological spaces. A \emph{homotopy} from
$X$ to $Y$ is a family of mappings $h_t:X\to Y$, $t\in I=[0,1]$
with the following property:
The mapping $H:X\times I\to Y$, $H(x,t)=h_t(x)$, is continuous. The set $X\times I$
has the product topology.
Two functions are called \DF{homotopic}, $f\cong g:X\to Y$ if a homotopy $h_t:X\to Y$
exists with $h_0=f$ and $h_1=g$. If $g$ is constant then $f$ is called \emph{nullhomotopic}.
A homotopy is called \DF{linear}, if it is linear in $t$.
\end{Def}
Just like in the definition of paths, the set $I$ does not need to be the set $[0,1]$ in the
discrete setting we are going to use arbitrary connected subsets of $\mathbb{Z}$ for instance
$\SET{0}{m}\subset\mathbb{N}$ with a fitting topology.
\begin{Def}
A topological space is called simply connected if any closed path is nullhomotopic.
\end{Def}
This means that we continuously contract every closed path into one point.
\begin{Lemma}\label{einfachverein}
If $(\mathcal{P},\mathscr{T})$ is a union of two open simply connected subspaces with contractible
intersection, then it is simply connected.\hspace{\stretch{1}}$\Box$
\end{Lemma}
\subsection{Alexandrov-Spaces}\label{section-alexandroff-space}
Every Alexandrov-space has an unique base that is given by the set of minimal
neighborhoods of all points in the base-set. The minimal neighborhoods are easily
identified as the intersections of all neighborhoods of a given point. Let $p$ be a
point in an Alexandrov-space $(\mathcal{P},\mathscr{T})$. We write $\mbox{\textnormal{U}}_\mathscr{T}(p)$ to denote its minimal
neighborhood. Analog we may find a minimal closed set containing a given point $p$.
We denote this set by $C_\mathscr{T}(p)$. To create an analogy to the graph-theoretic background
of most of this theory, we define
\begin{equation}
\mbox{\textnormal{A}}_\mathscr{T}(p):=(\mbox{\textnormal{U}}_\mathscr{T}(p)\cup\mbox{\textnormal{C}}_\mathscr{T}(p))\setminus\{p\}
\end{equation}
to be the \DF{adjacency} of the point $p$ in $(\mathcal{P},\mathscr{T})$. The set $\mbox{\textnormal{A}}_\mathscr{T}(p)$ can be made
to an Alexandrov-space in the subspace-topology.
Given a set $M\subset\mathcal{P}$ we may analog define the sets:
\begin{equation}\label{eq:closure1}
\mbox{\textnormal{U}}_\mathscr{T}(M) := \left\{p\in\mathcal{P}:\bigvee_{q\in M}p\in\mbox{\textnormal{U}}_\mathcal{T}`OP(q) \right\}
\end{equation}
\begin{equation}\label{eq:closure2}
\mbox{\textnormal{C}}_\mathscr{T}(M) := \left\{p\in\mathcal{P}:\bigvee_{q\in M}p\in\mbox{\textnormal{C}}_\mathscr{T}(q) \right\}
\end{equation}
\begin{Lemma}\label{huellenop}
The set functions $\mbox{\textnormal{U}}_\mathscr{T}$ and $\mbox{\textnormal{C}}_\mathscr{T}$ are closure operators, they satisfy:
\begin{enumerate}
\item $\mbox{\textnormal{U}}_\mathscr{T}(\emptyset)=\emptyset$.
\item $M\subset N\Rightarrow \mbox{\textnormal{U}}_\mathscr{T}(M)\subset \mbox{\textnormal{U}}_\mathscr{T}(N)$.
\item $\mbox{\textnormal{U}}_\mathscr{T}(\mbox{\textnormal{U}}_\mathscr{T}(M))=\mbox{\textnormal{U}}_\mathscr{T}(M)$.
\end{enumerate}
\end{Lemma}
\begin{Beweis}
The first property is trivial. To show the second one let $p\in\mbox{\textnormal{U}}_\mathscr{T}(M)$.
Therefore, it exists a $q\in M$ such that $p\in\mbox{\textnormal{U}}_\mathscr{T}(q)$.
By the precondition we have $q\in N$ and therefore $p\in\mbox{\textnormal{U}}_\mathscr{T}(N)$.
To prove property 3, let $p\in\mbox{\textnormal{U}}_\mathscr{T}(\mbox{\textnormal{U}}_\mathscr{T}(M))$, therefore, a $q$ exists in $\mbox{\textnormal{U}}_\mathscr{T}(M)$
such that $p\in\mbox{\textnormal{U}}_\mathscr{T}(q)$. If $q\in M$ holds, then holds $p\in\mbox{\textnormal{U}}_\mathscr{T}(M)$.
Otherwise, a $q'\in M$ exists such that $q$ is in $\mbox{\textnormal{U}}_\mathscr{T}(q')$.
By the property $T_0$ of an Alexandrov-space, the point $p$ has to be in $\mbox{\textnormal{U}}_\mathscr{T}(q')$
and therefore in $\mbox{\textnormal{U}}_\mathscr{T}(M)$. The other inclusion follows from 2.
\end{Beweis}
\begin{Lemma}\label{zus}
Let $(\mathcal{P},\mathscr{T})$ be an Alexandrov-space that contains one point $p$ such that the only open
neighborhood of $p$ is the set $\mathcal{P}$ itself. Then $(\mathcal{P},\mathscr{T})$ is contractible.
\end{Lemma}
\begin{Beweis}
We define a homotopy $F:\mathcal{P}\times I\to \mathcal{P}$ by $F(q,t)=q$
for $0\le t<1$ and $F(q,1)=p$ for each $q\in\mathcal{P}$.
We show, that $F$ is continuos. Let $M\subset \mathcal{P}$ be open.
Case 1: The point $p$ is in $M$. W.l.o.g. $M=\mathcal{P}$. Therefore, the set $F^{-1}(M)=\mathcal{P}\times I$
is open.
Case 2: The point $p$ is not in $M$. The the set $F^{-1}(M)=M\times [0,1)$ is open.
\end{Beweis}
\begin{Lemma}\label{alexzus}
Let $(\mathcal{P},\mathscr{T})$ be an Alexandrov-space and $p\in\mathcal{P}$, then the set $\mbox{\textnormal{U}}_\mathscr{T}(p)$ is
contractible. Therefore, the Alexandrov-space $(\mathcal{P},\mathscr{T})$ has a base of contractible
open sets. In particular, the set $(\mathcal{P},\mathscr{T})$ is local contractible.
\end{Lemma}
\begin{Beweis}
We utilize Lemma \ref{zus} together with $Y=\mbox{\textnormal{U}}(x)$ and $\omega=x$.
\end{Beweis}
It is possible to establish a notion of dimension in Alexandrov-spaces. It can also be
found in Evako et.al. \cite{evako}:
\begin{Def}\label{top_flaechen_def}
Let $(\mathcal{P},\mathscr{T})$ be a Alexandrov-space and $p\in\mathcal{P}$.
\begin{itemize}
\item $\mbox{\textnormal{dim}}_\mathcal{P}(p) := 0$, if $\mbox{\textnormal{U}}_\mathscr{T}(p)\setminus\{p\}=\emptyset$.
\item $\mbox{\textnormal{dim}}(\mathcal{P}) :=n$, if there is a point $p$ in $\mathcal{P}$ such that $\mbox{\textnormal{dim}}_\mathcal{P}(p)=n$
and for all $q\in\mathcal{P}$ exists a $k\le n$ with $\mbox{\textnormal{dim}}_\mathcal{P}(q)=k$.
\item $\mbox{\textnormal{dim}}_\mathcal{P}(p) := n+1$, if $\mbox{\textnormal{dim}}(\mbox{\textnormal{U}}_\mathscr{T}(p)\setminus\{p\})=n$.
The set $\mbox{\textnormal{U}}_\mathscr{T}(p)\setminus\{p\}$ has the subspace topology.
\item If no $k\in\mathbb{N}$ exists such that $\mbox{\textnormal{dim}}_\mathcal{P}(p)=k$ then define $\mbox{\textnormal{dim}}_\mathcal{P}(p)=\infty$.
\end{itemize}
\end{Def}
\begin{Def}
We call $(\mathcal{P},\mathscr{T})$ a $0$-surface, if $\mathcal{P}$ has two points and is disconnected under
$\mathscr{T}$.
The set $(\mathcal{P},\mathscr{T})$ is called $n$-surface for $n>0$, if $\mathcal{P}$ is connected under $\mathscr{T}$
and for all $p\in\mathcal{P}$ the set $\mbox{\textnormal{A}}_\mathscr{T}(p)$ is a $(n-1)$-surface.
A $n$-surface $(\mathcal{P},\mathscr{T})$ is called $n$-sphere, if $\mathcal{P}$ is finite and it is simply connected
for $n>1$.
\end{Def}
By Evako et.al.\cite{evako} gilt:
\begin{Satz}\label{alexandroffflaeche}
Let $(\mathcal{P},\mathscr{T})$ be a Alexandrov-space that is a $n$-surface for
$n>2$. Then, for any point $p\in\mathcal{P}$ holds, that $\mbox{\textnormal{A}}_\mathscr{T}(p)$
is simply connected. \hspace{\stretch{1}}$\Box$
\end{Satz}
\begin{Satz}\label{alexandrov-partial-order}
Every Alexandrov-space is a partial order and every partial order defines an
Alexandrov-space. \hspace{\stretch{1}}$\Box$
\end{Satz}
\subsection{The Khalimsky-Topology}\label{section-jordan-proof}
In this section we study an important Alexandrov-topologies. To define it we start with a topologization
of the set $\mathbb{Z}$ which we can interpret a a discrete line. What possibilities do we have to
define a non-trivial topology on this set such that it is connected?
One can see, that the sets
\begin{equation}
\mathcal{B} = \{\{x\}: x\in\mathbb{Z}, x\equiv 0(2)\} \cup
\{\{x-1,x,x+1\}:x\in\mathbb{Z},x\equiv 1(2)\}
\end{equation}
and
\begin{equation}
\mathcal{B'} = \{\{x\}: x\in\mathbb{Z}, x\equiv 1(2)\} \cup
\{\{x-1,x,x+1\}:x\in\mathbb{Z},x\equiv 0(2)\}
\end{equation}
are bases of topologies. They differ only by a translation. Therefore, it seams reasonable to
just choose one of them both. We will use the base $\mathcal{B}$ and denote its generated
topology by $\kappa$.
\begin{figure}[htb]
\begin{center}
\includegraphics{eps_files/basistopologie.eps}
\end{center}
\caption[Base of $(\mathbb{Z},\kappa)$]{\small{Figure (a) shows a section of $(\mathbb{Z},\kappa)$.
The base of the topology is represented by ellipses. The base of the topology may
also be depicted as a digraph. Figure (b) shows this.}}
\label{Pic:basistopologie}
\end{figure}
\begin{Lemma}
The Alexandrov-space $(\mathbb{Z},\kappa)$ is connected.\hspace{\stretch{1}}$\Box$
\end{Lemma}
To go from here to the higher-dimensional case, we may view $\mathbb{Z}^n$ as a $n$-fold
topological product of $\mathbb{Z}$. We denote the product topology
with $\kappa_n$. By all we know so far, it is clear, that $(\mathbb{Z}^n,\kappa_n)$ is
connected. We call this class of spaces \DF{Khalimsky-spaces} after E.~Khalimsky
\cite{khalimsky}.
\begin{Lemma}
The Alexandrov-space $(\mathbb{Z}^n,\kappa_n)$ is connected for all $n\ge 1$.
\end{Lemma}
\begin{Beweis}
This follows from Lemma \ref{prodzshg}.
\end{Beweis}
\begin{Satz}\label{svjb}
All Khalimsky-spaces $(\mathbb{Z}^n,\kappa_n)$, $n\ge2$ satisfy the separation theorem of
Jor\-dan-Brouwer.
\end{Satz}
\begin{Beweis}
The proof is easy if one uses the methods of algebraic topology, because
$(\mathbb{Z}^n,\kappa_n)$ is isomorphic to a cell-decomposition of $\mathbb{R}^n$:
\begin{equation}
\mathbb{R}^n =(\{\{i\}:i\in\mathbb{Z},i\equiv 0(2)\}\cup\{(i-1,i+1):i\in\mathbb{Z},i\equiv1(2)\})^n
\end{equation}
The set $(i-1,i+1)$ denotes the open real interval between the integers
$i-1$ and $i+1$.
Since the Theorem of Jordan-Brouwer is true for any $\mathbb{R}^n$, $n\ge2$, it has to
hold for $n$-dimensional Khalimsky-space.
We give another proof in section \ref{section-khalimsky-top}.
\end{Beweis}
\subsection{Adjacency Relations}
To establish structure on the points of the set $\mathbb{Z}^n$ we have to
define some kind of connectivity relation. This might be done in terms of a
(set-theoretic) topology as in the last section, or we may develop a graph-theoretic
framework as in the following part of the text.
\begin{Def}\label{adjazenzrelation}
Given a set $\mathcal{P}$, a relation $\alpha\subset\mathcal{P}\times\mathcal{P}$ is called \DF{adjacency}
if it has the following properties:
\begin{enumerate}
\item $\alpha$ is finitary: $\forall p\in\mathcal{P}:|\alpha(p)|<\infty$.
\item $\mathcal{P}$ is connected under $\alpha$.
\item Every finite subset of $\mathcal{P}$ has at most one infinite connected component
as complement.
\end{enumerate}
\end{Def}
A set $M\subset\mathcal{P}$ is called \DF{connected} if for any two points
$p,q$ in $M$ exist points $p_0,\ldots,p_m$ and a positive integer $m$ such that
$p_0=p$, $p_m=q$ and $p_{i+1}\in A(p_i)$ for all
$i\in\{0,\ldots,m-1\}$. Compare this definition to the topological one we gave above.
The property 3 of an adjacency-relation is in $\mathbb{Z}^n$ for $n\ge 2$ always satisfied.
In the text we will consider pairs $(\alpha,\beta)$ of adjacencies
on the set $\mathbb{Z}^n$. In this pair $\alpha$ represents the adjacency
on a set $M\subset\mathbb{Z}^n$, while $\beta$ represents the adjacency on
$\COMP{M}=\mathbb{Z}^n\setminus M$.
Let $\mathcal{T}$ be the set of all translations on the set $\mathbb{Z}^n$.
The generators $\tau_1,\ldots,\tau_n\in\mathcal{T}$ of $\mathbb{Z}^n$
induce a adjacency $\pi$ in a natural way:
\begin{Def} Two points $p,q$ of $\mathbb{Z}^n$ are called \DF{proto-adjacent}, in terms
$p\in\pi(q)$, if there exists a $i\in\{1,\ldots,n\}$ such that
$p=\tau_i(q)$ or $p=\tau_i^{-1}(q)$.
\end{Def}
We can view the generators of $\mathbb{Z}^n$ a the standard base of $\mathbb{R}^n$.
Another important adjacency on $\mathbb{Z}^n$ is \DF{$\omega$}.
\begin{equation}
\omega(p):=\{q\in\mathbb{Z}: |p_i-q_i|\le 1, 0\le i\le n\}
\end{equation}
In the rest of the text let $\alpha$ and $\beta$ be two adjacencies on
$\mathbb{Z}^n$ such that for any $p\in\mathbb{Z}^n$ holds
\begin{equation}
\pi(p)\subset\alpha(p),\beta(p)\subset\omega(p)\enspace.
\end{equation}
\begin{Lemma}
The set $\mathbb{Z}^n$ is connected under $\pi$.\hspace{\stretch{1}}$\Box$
\end{Lemma}
\section{Digital Manifolds}
If we want to talk about $(n-1)$-Manifolds in $\mathbb{Z}^n$ we have to give a proper definition.
Unfortunatly, all the definitions known to the author from the literature are not usable
in terms of generalization to higher dimension or for the unification of the topological
and graph-theoretic approach. So it is necessary, to give a new definition that
satisfies this two criteria. This is don in \cite{paper01}. The new definition is
manly based on the so called \emph{separation property}. It gives a description on
how a discrete $(n-1)$-manifold should look like locally.
\subsection{The Separation Property}\label{Sec:trennungseigenschaft}
We call the set
\begin{equation}
C^k=\{0,1\}^k\times\{0\}^{n-k}\subset\mathbb{Z}^n
\end{equation}
the $k$-dimensional standard \DF{cube} in $\mathbb{Z}^n$.
The set $C^k$ can be embedded in ${n}\choose{k}$ different ways in $C^n$.
A general $k$-cube in $\mathbb{Z}^n$ is defined by a translation of a standard cube.
Indeed, we can construct any $k$-cube $C$ from one point $p$ with $k$ generators
in the following way:
\begin{equation}
C = \{\tau_1^{e_1}\cdot\tau_k^{e_k}(p): e_i\in\{0,1\}, i=1,\ldots,k \}
\end{equation}
The dimension of $C'$ is then $k+l$. We use this construction in the next definition.
\begin{Def}\label{trenndef}
Let $M\subset \mathbb{Z}^n$, $n \ge 2$ and $C$ be a $k$-cube,
$2\le k\le n$. The complement of $M$ is in $C$ \DF{not separated} by $M$ under
the pair $(\alpha,\beta)$, if
for every $\alpha$-component $M'$ of $C\cap M$ and every
$(k-2)$-subcube $C^*$ of $C$ the following is true:
If $C^*$ is such that $C^*\cap M'\neq\emptyset$ has maximal cardinality among
all sets of this form, and the sets $\tau_1(C^*)\setminus M$ and
$\tau_2(C^*)\setminus M$ are both nonempty and lie in one common $\beta$-component
of $\COMP{M}$, then holds
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C^*)\cap M') \subset
\tau_1^{-1}(\tau_1(C^*)\cap M') \cap
\tau_2^{-1}(\tau_2(C^*)\cap M')
\enspace .
\end{equation}
\end{Def}
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=1.0]{eps_files/trennungseigenschaft1.eps}
\end{center}
\caption[The separation property]{\small{$C^*$ has an intersection
of maximal cardinality with the $\alpha$-component
$M'$. The sets $\tau_1(C^*)\setminus M'$ and $\tau_2(C^*)\setminus M'$ are
nonempty and belong to a $\beta$-component of
$\COMP{M}$. Since $\tau_1\tau_2(C^*)\cap M' = \emptyset$,
the property of definition \ref{trenndef} is satisfied for this $C^*$.
But the set $M'$ separates $\COMP{M}$ in the cube $C^*$. Why?}}
\label{Pic:trennungseigenschaft1}
\end{figure}
In the following, we only consider the case when $C\cap M$ has at most one
$\alpha$-component. This can be justified by viewing any other
$\alpha$-component besides the one considered as part of the
background, since there is no $\alpha$-connection anyway.
This property also gets important if we study the construction of the
simplicial complex.
A set $M$ has the \DF{separation property} under a pair $(\alpha,\beta)$,
if for every $k$-cube $C$, $2\le k\le n$ as in the definition \ref{trenndef}
the set $\COMP{M}$ is in $C$ not separated by $M$
The meaning of the separation property is depicted in the figure
\ref{Pic:trennungseigenschaft}.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=1.0]{eps_files/trennungseigenschaft.eps}
\end{center}
\caption[The separation property 2]{\small{The black points represent the set $M$ in the
given 3-cubes. The white points represent the complement of $M$. In the cases (a) to (c)
the complement, which is connected, is separated by $M$. This separation is depicted
by the gray plane spanned by $C^*$ and $\tau_1\tau_2(C^*)$.
In Figure (d) occurs no separation,
since the only choice for $C^*$ would be a 1-cube, that contains only black
points.
}}
\label{Pic:trennungseigenschaft}
\end{figure}
\begin{Def}\label{n-1-mannigfaltigkeit}
An $\alpha$-connected set $M\subset\mathbb{Z}^n$, for $n\ge 2$, is a
\DF{(digital) $(n-1)$-manifold} under the pair
$(\alpha,\beta)$, if the following properties hold:
\begin{enumerate}
\item\label{cubeconnection} In any $n$-cube $C$ the set $C\cap M$
is $\alpha$-connected.
\item\label{twocomponents} For every $p\in M$ the set
$\omega(p)\setminus M$ has exactly two $\beta$-components
$C_p$ and $D_p$.
\item\label{componentunity} For every $p\in M$ and every
$q\in\alpha(p)\cap M$ the point $q$ is $\beta$-adjacent to $C_p$ and $D_p$.
\item\label{separationproperty} $M$ has the separation property.
\end{enumerate}
\end{Def}
How should a $(n-1)$-manifold look like globally in general? We do not know. But we might say,
that a single point in $\mathbb{Z}^n$ might be considered as the inside of some object, i.e.
that it might be separated by the other points. The way to do this is to require the set
of neighbors of a point to be a $(n-1)$-manifold. This justifies the following:
\begin{Def}
A pair $(\alpha,\beta)$ of adjacency relations on $\mathbb{Z}^n$ is a
\DF{separating pair} if for all $p\in\mathbb{Z}^n$ the set $\beta(p)$ is a
$(n-1)$-manifold under $\alpha$.
\end{Def}
\subsection{Double Points}
\begin{Def}\label{doppelpunkt}
A point $p\in\beta(z)$ $z\in\mathbb{Z}^n$ is a \DF{double point} under the
pair $(\alpha,\beta)$, if there exist points
$q\in\pi(z)\cap\alpha(p)$ and $r\in\beta(z)\cap\pi(p)$
and a simple\footnote{A translation $\tau$ is called \DF{simple} if no other translation
$\sigma$ exists with $\sigma^n=\tau$, $n\in\mathbb{Z}, |n|\neq 1$. }
translation $\tau\in\mathcal{T}$ with $\tau(p)=q$, $\tau(r)=z$ and $q\in\alpha(r)$.
\end{Def}
This concept is the key to a local characterization of the good pairs
$(\alpha,\beta)$. Without it, one could not consistently define topological
invariants like the Euler-char\-acter\-istic. It means that an edge
between points in a set $M$ can be crossed by an edge between points
of its complement and these four points lie in a square defined by the
corresponding adjacencies. This crossing can be seen as a double point,
belonging both to the foreground \emph{and} to the background. Also,
mention the close relationship to the separation property, which is a
more general concept of similar interpretation. For further insight, refer
to the text \cite{paper01}.
\begin{figure}[htb]
\begin{center}
\includegraphics{eps_files/doppelpunkt.eps}
\end{center}
\caption[A double point]{\small{A double point $p$. The fat edges
represent the $\alpha$-adjacency. Only the relevant edges have been drawn for
clarity. The dotted edge represents the $\beta$-adjacency of $p$ and $z$. The
black points are the $\beta$-neighbors of $z$.}}
\label{Pic:doppelpunkt}
\end{figure}
\begin{Def}\label{goodpair}
A separating pair of adjacencies $(\alpha,\beta)$ in
$\mathbb{Z}^n$ is a \DF{good pair}, if for every $p\in\mathbb{Z}^n$ the set
$\beta(p)$ contains no double points.
\end{Def}
\section{Good Pairs of Adjacency Relations}\label{chapter_6}
\subsection{Cubical Adjacencies}\label{kubische_adjazenzen}
We will study adjacencies in the sense of the gridcube-model. This is a common model
in computer graphics literature and has nothing to do with the $n$-cubes we talked about
earlier. We use this model here to make it easy to study the adjacency relations in this
section. For more on this topic refer to the Book of Rosenfeld and Klette \cite{klette}
We identify the points of $\mathbb{Z}^n$ with $n$-dimensional unit-cubes with barycenters in the
points of the lattice $\mathbb{Z}^n$. The cube $W$ that represents the point $0\in\mathbb{Z}^n$ can be
expressed in euclidean space as $[-\frac{1}{2},\frac{1}{2}]^n$. Those gridcubes may
be interpreted as union of (polytopal) faces of different dimension. Any of its faces
is again a gridcube, only with a lower dimension. Take, for instance, a 3-dimensional
gridcube $[-\frac{1}{2},\frac{1}{2}]^3$. It has, among others, the 0-dimensional face
$(\frac{1}{2},\frac{1}{2})$, the 1-dimensional face
$[(-\frac{1}{2},\frac{1}{2}),(\frac{1}{2},\frac{1}{2})]$
and the 2-dimenional face with the vertices
$(-\frac{1}{2},\frac{1}{2}), (\frac{1}{2},\frac{1}{2}),
(-\frac{1}{2},-\frac{1}{2})$ and $(\frac{1}{2},-\frac{1}{2})$.
Two given gridcubes may share a $k$-dimensional face for $0\le k<n$. This
$k$-face is just the intersection of both of them. So we might say that the elements
$(0,\ldots,0)$ and $(1,\ldots,1)$ of $\mathbb{Z}^n$ intersect in a common vertex (0-face) with
the coordinates $(\frac{1}{2},\ldots\frac{1}{2})$. However, the elements $(0,\ldots,0)$
and $(1,0,\ldots,0)$ share a common $(n-1)$-face.
In the rest of the text we will no longer make the gridcube model explicit. It just serves
as an introduction to visualize the concepts that we use to analyze the discrete
geometry even in higher dimensions\footnote{I find it a lot easier to imagine a four-dimensional
cube, than a four-dimensional grid...}.
\begin{Def} Two points $p,q\in\mathbb{Z}^n$ are called \DF{$k$-adjacent} for $0\le
k<n$, denoted by $p\in\alpha_k(q)$, if their corresponding gridcubes share a common
$k$-face. We call this adjacencies \DF{cubical}.
\end{Def}
Clearly, this kind of relation we just defined is an adjacency-relation in the sense of
definition \ref{adjazenzrelation}:
\begin{Lemma}
The relation $\alpha_k$ is an adjacency-relation on $\mathbb{Z}^n$
for every $n\ge 2$ and all integers $k$ between $0$ and $n-1$.
\end{Lemma}
\begin{Beweis}
First, we have to check that for any $p\in\mathbb{Z}^n$ the set $\alpha_k(p)$ has only
finite cardinality. It is easy to check, that $\alpha_0(p)$ is just $\omega(p)$
as defined earlier and every $\alpha_k(p)$ for $0\le k<n$ is a subset of $\omega(p)$.
Since $\omega(p)$ has $3^n-1$ Elements in $\mathbb{Z}^n$, the relations $\alpha_k$ must be
finitary.
To see that $\mathbb{Z}^n$ is connected under any $\alpha_k$, $0\le k<n$, we observe that
$\alpha_{n-1}$ is just another interpretation for the relation $\pi$ defined earlier.
Since $\mathbb{Z}^n$ is $\pi$-connected as proven in \cite{paper01} and every $\alpha_k$ is a
superset of $\alpha_{n-1}$, we conclude that $\mathbb{Z}^n$ is $\alpha_k$-connected.
The last property is in $\mathbb{Z}^n$ with $n\ge 2$ trivially satisfied.
\end{Beweis}
\begin{Lemma}\label{kuben_representation}
The cubical adjacency $\alpha_k(x_1,\ldots,x_n)$ may be represented in $\mathbb{Z}^n$ as the
set:
\begin{equation}
\left\{(y_1,\ldots,y_n)\in\mathbb{Z}^n:
\max_{i=1,\ldots,n}\{|x_i-y_i|\}=1,1\le\sum_{i=1}^n|x_i-y_i|\le n-k\right\}
\end{equation}
\end{Lemma}
\begin{Beweis}
Let $p$ and $q$ be two points of $\mathbb{Z}^n$ such that $p\in\alpha_k(q)$.
This means, the gridcubes corresponding to $p$ and $q$ share a common $k$-face.
Their distance in the maximum-metric may not be greater than 1. Furthermore,
$p$ and $q$ may not share a single common $l$-face for $0\le l<k$. That means, all of that
$l$-faces must be faces of common $k$-faces. Therefore, the two points may not have
more than $k$ coordinates in common.
\end{Beweis}
\begin{Lemma}
Let $\alpha$ be a cubical adjacency on $\mathbb{Z}^n$. It holds:
\begin{enumerate}
\item $\alpha$ is invariant under translations
\item $\alpha$ is invariant under permutations of coordinates.
\end{enumerate}
\end{Lemma}
\begin{Beweis}
Let $\tau$ be any translation on $\mathbb{Z}^n$. We need to show $\tau(\alpha(p))=\alpha(\tau(p))$
for any $p\in\mathbb{Z}^n$.
From the representation of $\alpha(p)$ we may deduce:
\begin{eqnarray}
\tau(\alpha(p)) &=& \tau(\{q\in\mathbb{Z}^n:q\in\alpha(p)\} \\
&=& \{\tau(q):q\in\mathbb{Z}^n, q\in\alpha(p)\} \\
&=& \{q'\in\mathbb{Z}^n: q'\in\alpha(\tau(p)) \} \\
&=& \alpha(\tau(p))
\end{eqnarray}
The proof of the second part is analog.
\end{Beweis}
What is the structure of the cubical adjacencies in $\mathbb{Z}^n$? We take a closer look
at $n$-dimensional cubes.
\begin{Lemma}
The number of $k$-faces of a $n$-dimensional cube is
\begin{equation}
{{n}\choose{k}}\cdot 2^{n-k}\enspace.
\end{equation}
\end{Lemma}
\begin{Beweis}
We use induction on the dimension $n$ of the cube.
For $n=0$ we observe, that a 0-dimensional cube is just a point and has only one 0-face.
Therefore, the induction base is correct.
In the case $n>0$, we notice that a $n$-dimensional cube may be created from a
$(n-1)$-dimensional one by doubling the cube and inserting a $k$-face for every
$(k-1)$-face in the original cube. Therefore, we get by induction hypothesis and Pascals
Theorem:
\begin{eqnarray}
2{{n-1}\choose{k}}\cdot2^{n-1-k}+{{n-1}\choose{k-1}}\cdot2^{n-1-(k-1)}
&=& \left[{{n-1}\choose{k}}+{{n-1}\choose{k-1}}\right]\cdot2^{n-k} \\
&=& {n \choose k}\cdot2^{n-k}
\end{eqnarray}
This proves the Lemma.
\end{Beweis}
\begin{Lemma}
For every $p\in\mathbb{Z}^n$, the number of $k$-neighbors is
\begin{equation}
|\alpha_k(p)| = \sum_{i=k}^n{{n}\choose{i}}\cdot2^{n-i}\enspace.
\end{equation}
\end{Lemma}
\begin{Beweis}
Obviously, any $l$-face $\sigma$ of a cube contains at least one $k$-face
$\tau$ for $0\le k\le l \le n$. Therefore, $k$-adjacent cubes exist, that are also $l$-adjacent.
Since that are those, that share more than one common $k$-face, the set
$\alpha_k(p)$ for $p\in\mathbb{Z}^n$ may be decomposed into the following disjoint sets:
\begin{equation}
\begin{split}
\alpha_k(p) &=\{q\in\mathbb{Z}^n:\mbox{$p,q$ have at most one $k$-face in common }\} \\
&\cup\,\,\{q\in\mathbb{Z}^n:\mbox{$p,q$ have at most one $(k+1)$-face in common }\}\\
&\,\,\,\,\vdots \\
&\cup\,\,\{q\in\mathbb{Z}^n:\mbox{$p,q$ have at most one $(n-1)$-face in common }\}
\end{split}
\end{equation}
By adding the cardinalities of these sets, which we can easily compute with
the last Lemma we get the result
$\alpha_k(p)=\sum_{i=1}^n{{n}\choose{i}}2^{n-i}$.
This proves the Lemma.
\end{Beweis}
By this technique we get as examples of cubical adjacencies in $\mathbb{Z}^2$ the known
4- and 8-adjacencies, in $\mathbb{Z}^3$ the 6-, 18- and 26-adjacencies and in $\mathbb{Z}^4$ the
8-, 32-, 64- and 80-adjacencies.
\subsection{Good Pairs of Cubical Adjacencies}
In this section we will study, how we have to choose two cubical adjacencies to get to a
good pair. We first will see, that it does not matter at which point of $\mathbb{Z}^n$ we study
the adjacency, since the neighborhoods of all points look the same.
\begin{Lemma}
Let $\alpha$ be a cubical adjacency in $\mathbb{Z}^n$. For any $p\in\mathbb{Z}^n$ the set
$\alpha(p)$ is graph-theoretical isomorphic to $\alpha(0)$.
\end{Lemma}
\begin{Beweis}
This follows from the invariance under translations and the symmetry of the cubical
adjacencies.
\end{Beweis}
\begin{Lemma}\label{kleiner-zushang}
Let $M\subset\mathbb{Z}^n$ be $\alpha_k$-connected. Then $M$ is also
$\alpha_l$-connected for $0\le l<k\le n-1$.
\end{Lemma}
\begin{Beweis}
Let $M$ be $\alpha_k$-connected. Thus, we have for any two $p,q\in M$ a path
$p=p^{(0)},\ldots,p^{(a)}=q$ such that $p^{(i)}\in M$ for
$i\in\{0,\ldots,a\}$ and $p^{(i-1)}\in\alpha_k(p^{(i)})$ for
$i\in\{1,\ldots,a\}$. By definition of $\alpha_k$, Lemma \ref{kuben_representation} and
$l<k$ holds for $p^{(i-1)}$ and $p^{(i)}$:
$|p^{(i)}_j-p^{(i-1)}_j|\le 1$ and
\begin{equation}
1\le\sum_{j=1}^n|p^{(i)}_j-p^{(i-1)}_j|\le n-k < n-l\enspace.
\end{equation}
Therefore, we have $p^{(i-1)}\in\alpha_k(p^{(i)})$ for $i\in\{1,\ldots,a\}$
and the path $p=p^{(0)},\ldots,p^{(a)}=q$ is also a $\alpha_l$-path.
\end{Beweis}
The next Lemmata help us understand, which adjacencies may be used as good pair.
\begin{Lemma}
Let $(\alpha_l,\alpha_k)$ be a pair of cubical adjacencies on $\mathbb{Z}^n$,
$n\ge 2$. For any $n$-cube $C$ as in section \ref{Sec:trennungseigenschaft}, the set
$C\cap\alpha_k(0)$ is connected under $\alpha_l$ if the following holds:
\begin{enumerate}
\item $0\le k\le n-2$ and $0\le l\le n-1$, or
\item $k=n-1$ and $0\le l\le n-2$.
\end{enumerate}
\end{Lemma}
\begin{Beweis}
1. We use Lemma \ref{kleiner-zushang} and prove the proposition for $l=n-1$
Let $C'$ be any subcube of $C$, that does not contain the point 0.
We first show that $C'\cap M$ is $\alpha_l$-connected. Suppose w.l.o.g. that
the point $p=(1,0,\ldots,0)$ is in $C'$ and choose any other point $r\in C'\setminus M$.
The point $r$ then has the form $r=(1,r_2,\ldots,r_n)$ with
\begin{equation}
\max_{i=1,\ldots,n}|r_i|=1 \mbox{ and } 1\le 1 +
\sum_{i=2}^n|r_i|\le n-k\enspace.
\end{equation}
We select the smallest index $i\in\{2,\ldots, n\}$ such that $r_i\neq 0$ and define
\begin{equation}
r'=(1,r_2,\ldots,r_{i-1},0,r_{i+1},\ldots,r_n)\enspace.
\end{equation}
The point $r'$ is in $\alpha_{n-1}(r)$:
\begin{equation}
\max_{i=1,\ldots,n}|r'_i|=1 \mbox{ and } 1\le
\sum_{i=1}^n|r'_i-r_i|=|r'_i-r_i|=1\le n-(n-1)\enspace.
\end{equation}
By iterating this process we get an $\alpha_{n-1}$-path from $r$ to $p$.
Let now be $C'$ and $C''$ be two different $(n-1)$-cubes. We may suppose w.l.o.g.
that $p=(1,0,\ldots,0)\in C'$ and $q=(0,1,0,\ldots,0)\in C''$.
The two cubes contain a common point $t=(1,1,0,\ldots,0)$ in $M$ since this point is
$\alpha_{n-1}$-adjacent to $p$ and $q$ and it is in $\alpha_k(0)$ for
$0\le k\le n-2$:
\begin{equation}
\max_{i=1,\ldots,n}|t_i|=1 \mbox{ and } \sum_{i=1}^n|t_i| = 2\le n-k\enspace.
\end{equation}
Therefore, the set $C\cap\alpha_k(0)$ is $\alpha_{n-1}$-connected.
2. We show, that $C\cap\alpha_{n-1}$ is connected under $\alpha_{n-2}$.
By Lemma \ref{kleiner-zushang} this is enough.
The set $C\cap\alpha_{n-1}$ contains all points $p^{(i)}=(p_1,\ldots,p_n)$, such that
exactly one $i\in\{1,\ldots,n\}$ exists with $p^{(i)}_i\neq 0$ and
$|p^{(i)}_i|=1$. Let $p^{(i)}$ and $p^{(j)}$ be two such points with $i\neq j$.
We have
\begin{equation}
\max_{m=1,\ldots,n}|p^{(i)}_m-p^{(j)}_m|=1 \mbox{ and }
\sum_{m=1}^n|p^{(i)}_m-p^{(j)}_m|=2\le n-(n-2)\enspace.
\end{equation}
Therefore, $p^{(i)}$ and $p^{(j)}$ are $\alpha_{n-2}$-connected.
\end{Beweis}
\begin{Folg}\label{kub-surf}
Given a pair $(\alpha_l,\alpha_k)$, then $\alpha_k(0)$ is
$\alpha_l$-connected, if the following holds:
\begin{enumerate}
\item $0\le k\le n-2$ and $0\le l\le n-1$ or
\item $k=n-1$ and $0\le l\le n-2$.
\end{enumerate}
\end{Folg}
\begin{Beweis}
This follows from the configuration of the $n$-cubes in $\omega(0)$ and the distribution
of the $\pi$-neighbors of 0 in those $n$-cubes
\end{Beweis}
\begin{Lemma}
Let $(\alpha_l,\alpha_k)$ be a pair of cubical adjacencies on $\mathbb{Z}^n$ with $n\ge 2$.
Then the set $\omega(p)\setminus\alpha_k(0)$ has exactly two $\alpha_k$-components
for all $p\in\alpha_{k}{0}$.
\end{Lemma}
\begin{Beweis}
Obviously, 0 is in $\omega(p)$ for any $p\in\alpha_k(0)$ and it has no other
$\alpha_k$-neighbors in $\alpha_k(0)\setminus\omega(p)$.
We choose any point $p$ in $\alpha_k(0)$. Then, $\omega(p)$ contains points $s$ with
$\max_{i=1,\ldots,n}|s_i|=2$. Those are not contained in in $\alpha_k(0)$ and form a
$\pi$-connected set. Therefore they are also $\alpha_k$-connected.
Define the set:
\begin{equation}
\omega(p)_i:=\{s\in\omega(p): |s_i|=2\}\enspace.
\end{equation}
W.l.o.g. we consider $\omega(p)_1$ that contains the point $p'=(2,p_2,\ldots,p_n)$.
It is easy to see, that either the point $p''=(-2,p_2,\ldots,p_n)$ or the point
$p'$ is in $\omega(p)_1$.
Let $s=(2,s_2,\ldots,s_n)$ be any point in $\omega(p)_i$.
We construct a $\pi$-path from $s$ to $p'$ by defining the point
\begin{equation}
s'=(2,\ldots,p_2,\ldots,p_{i-1},p_i,s_{i+1},\ldots,s_n)
\end{equation}
with the smallest index $i\in\{2,\ldots,n\}$ such that $s_i\neq p_i$.
The point $s'$ is a $\pi$-neighbor of $s$ and after a finite number of iterations
we have the $\pi$-path from $s$ to $p'$.
The sets $\omega(p)_i$ and $\omega(p)_j$ contain the points
\begin{equation}
(p_1,\ldots,p_{i-1},2,p_{i+1},\ldots,p_n)
\end{equation}
and
\begin{equation}
(p_1,\ldots,p_{j-1},2,p_{j+1},\ldots,p_n)\enspace,
\end{equation}
respectively. In both sets the point
\begin{equation}
t=(p_1,\ldots,p_{i-1},2,p_{i+1},\ldots,p_{j-1},2,p_{j+1},\ldots,p_n)
\end{equation}
is contained and therefore, the sets are $\pi$-connected.
It remains to show, that points in $\omega(0)\setminus\alpha_k(0)\cap\omega(p)$
are $\pi$-adjacent to one of the $\omega(p)_i$. Let $i\in\{1,\ldots,n\}$, such that
$s_i=p_i\neq 0$.
In the case $p_i>0$, then we have
\begin{equation}
s'=(s_0,\ldots,s_{i-1},s_i+1,s_{i+1},\ldots,s_n)\in\omega(p)_i
\end{equation}
and in the case $p_i<0$, it holds
\begin{equation}
s'=(s_0,\ldots,s_{i-1},s_i-1,s_{i+1},\ldots,s_n)\in \omega(p)_i\enspace.
\end{equation}
Finally, we have to observe the case of the point $s$ with $0=s_i\neq p_i$.
Then, the point
\begin{equation}
s'=(s_1,\ldots,s_{i-1},p_i,s_{i+1}\ldots,s_n)
\end{equation}
is a $\pi$-neighbor
of $s$, that is not in $\alpha_k(0)$.
This follows from
\begin{equation}
n-k<\sum_{j=1}^n|s_j|<\sum_{j=1}^{i-1}|s_j|+|p_i|+\sum_{j+1}^n|s_j|
=\sum_{j=1}^n|s_j|+1\enspace.
\end{equation}
Thus, in the set $\alpha_k(0)\setminus\omega(p)$, there is only one $\alpha_k$-component
different from 0.
\end{Beweis}
\begin{Lemma}\label{zahl-der-komponenten-von-alpha-k}
Let $(\alpha_l,\alpha_k)$ be a pair of cubical adjacencies on
$\mathbb{Z}^n$ with $n\ge 2$. For any $p\in\alpha_k(0)$ any point
$q\in\alpha_l(p)\cup\alpha_k(0)$ is $\alpha_k$-adjacent to both of the
$\alpha_k$-components of $\omega(p)\setminus\alpha_k(0)$, if the following
holds:
\begin{enumerate}
\item $1\le k\le n-1$ and $0\le l\le n-1$
\item $0\le k\le n-2$ and $l=n-1$
\end{enumerate}
\end{Lemma}
\begin{Beweis}
1. Let $p\in\alpha_k(0)$ and $q\in\alpha_l\cap\alpha_k(0)$ be arbitrary chosen.
Since $q\in\alpha_k(0)$, the point $q$ is $\alpha_k$-adjacent to the
$\alpha_k$-component $\{0\}$.
We define the set
\begin{equation}
I(p,q):=\{1\le i\le n:p_i=q_i=0\}\enspace.
\end{equation}
This set is non-empty since $1\le k\le n-1$. Let $q'=(q'_1,\ldots,q'_n)$ be the point with
the following coordinates
\begin{equation}
q'_i=\left\{\begin{array}{cl}
q_i&\mbox{if }q_i\neq 0\\
p_i&\mbox{if }q_i=0\mbox{ and }p_i\neq 0\\
\end{array}\right.
\end{equation}
The remaining $q'_i$ for $i\in I(p,q)$ will be assigned with the values $\pm1$ and 0
such that exactly $n-k+1$ of $n$ coordinates are different from 0.
The point $q'$ is not contained in $\alpha_k(0)$, since
\begin{equation}
\sum_{i=1}^n|q_i'|=n-k+1>n-k\enspace.
\end{equation}
Because of $|q_i'-p_i|\le 1$, the point $q'$ must be in $\omega(p)$. Therefore, the point
$q'$ is in $\omega(p)\setminus(\alpha_k(0)\cup\{0\})$.
Since $q$ is not 0, we have
\begin{equation}
\sum_{i=1}^n|q_i'-q_i|=\sum_{i=1}^n|q_i'|-\sum_{i=1}^n|q_i|\le
n-k\enspace.
\end{equation}
Therefore, the point $q'$ is in $\alpha_k(q)$ and the set
$\omega(p)\setminus(\alpha_k(0)\cup\{0\})$ is
$\alpha_k$-adjacent to $q$.
For $k=n-1$ and $l=n-1$ the set $\alpha_l(p)\cap\alpha_k(0)$ is empty.
Thus the proposition is true.
We may not choose $k$ as 0 for $0\le l\le n-2$, as the following example shows:
Let $p\in\pi(0)$ and
$q\in\alpha_l(p)\cap\pi(0)$, then
\begin{equation}
\sum_{i=1}^n|p_i-q_i|=2\le n-l\enspace,
\end{equation}
and thus
$q\in\alpha_l(p)$. The point $q$ has no $\alpha_0$-neighbors
in $\omega(p)\setminus\alpha_0(0)$, because of $\alpha_0=\omega$ and
because $r_i=\pm2$ for $p_i=\pm1$, hold for all $r\in\omega(p)\setminus(\omega(0)\cup\{0\}$.
2. Let $p$ be a point in $\alpha_k(0)$ and let
$q$ be any point in $\alpha_{n-1}(p)\cap\alpha_k(0)$. Obviously, the point
$q$ is $\alpha_k$-adjacent to $\{0\}$.
We choose
\begin{equation}
i\in\{1\le i\le n: p_i\neq 0\mbox{ and }
q_i\neq 0\}\enspace.
\end{equation}
This is a non-empty set, because the points $q$ and $p$ coincide in at least one
non-zero coordinate, since $q\in\alpha_{n-1}(p)\cap\gamma_k(0)$.
We define
\begin{equation}
q'=\left\{\begin{array}{cl}
(q_1,\ldots,q_{i-1},q_i+1,q_{i+1},\ldots,q_n)&\mbox{if } p_i=1\\
(q_1,\ldots,q_{i-1},q_i-1,q_{i+1},\ldots,q_n)&\mbox{if } p_i=-1\\
\end{array}\right.
\end{equation}
We have $|q_j'-p_j|\le 1$ for all $i\in\{1,\ldots,n\}$. Therefore, the point
$q'$ is in $\omega(p)$. Thus, $|q_i|=2$, otherwise $|q_i-p_i|$ would not be
smaller or equal to 1.
We conclude that $q'$ is no point in $\alpha_k(0)$ and $q'\neq 0$.
The point $q'$ is therefore a member of $\omega(p)\setminus(\alpha_k(0)\cup\{0\})$
and it is a $\pi$-neighbor of $q$. Which means it is an
$\alpha_k$-neighbor, too.
Thus, the point $q$ is $\alpha_k$-adjacent to
$\omega(p)\setminus(\alpha_k(0)\cup\{0\})$.
\end{Beweis}
\begin{Lemma}
Let $(\alpha_l,\alpha_k)$ be a pair of cubical adjacencies. Then the set
$\alpha_k(0)$ satisfies the separation property under this pair.
\end{Lemma}
\begin{Beweis}
Instead of $\alpha_k(0)$, we consider the set
$\overline\alpha_k(0)=\alpha_k(0)\cup\{0\}$. We may do so, because the point $0$
is a different $\alpha_k$-component of $\COMP{\alpha_k(0)}$. It is not separable
and therefore has no influence of the separability of the other points. If we know
whether $\overline\alpha(0)$ has the separation property,
then we also know that $\alpha_k(0)$ has it too.
Let $C$ be a $m$-cube, $2\le m\le n$, containing a point of $\overline\alpha_k(0)$ and
let $C^*$ be a $(m-2)$-subcube of $C$ that has a maximal intersection with
$\overline\alpha_k(0)$. There exist two translations $\tau_1$ and $\tau_2$ such that we may
decompose $C$ in the following way:
\begin{equation}
C=C^*\cup\tau_1(C^*)\cup\tau_2(C^*)\cup\tau_1\tau_2(C^*)\enspace.
\end{equation}
Case 1: The translated cubes $\tau_i(C^*)$, $i=1,2$ and $\tau_1\tau_2(C^*)$ each
contain a point $p$ such that $\max_{i=1,\ldots,n}|p_i|=2$.
Then, one can easily see that $\tau_1\tau_2(C^*)$ is fully in $\COMP{M}$.
And so we have
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C^*)\cap M)\subset
\tau_1^{-1}(\tau_1(C^*)\cap M)\cap\tau_2^{-1}(\tau_2(C^*)\cap M)
\subset C^*\cap M\enspace.
\end{equation}
This is true, especially if
$\tau_i(C^*)\setminus\overline\alpha_k(0)\neq\emptyset$, $i=1,2$ and both of the
set are $\alpha_k$-connected.
Case 2: The cube $C$ contains no point $p$ such that
$\max_{i=1,\ldots,n}|p_i|=2$. Let $q$ be the point in $C^*$ that satisfies
$\sum_{i=1}^n|q_i|=x$ and $0\le x\le n-k$ be minimal in $C$.
It is sufficient to claim this minimality as the following consideration shows:
We have:
\begin{equation}
C^*\cap M=\left\{p:\sum_{i=1}^n|p-i|\le \min(m-2,n-k)-x \right\}
\end{equation}
\begin{equation}
\tau_{1,2}C^*\cap M=\left\{p:\sum_{i=1}^n|p-i|\le
\min(m-3,n-k-1)-x \right\}
\end{equation}
\begin{equation}
\tau_1\tau_2(C^*)\cap M=\left\{p:\sum_{i=1}^n|p-i|\le
\min(m-4,n-k-2)-x \right\}
\end{equation}
The cube $C^*$ has always a maximal number of points in $\overline\alpha_k(0)$.
If $\tau_{1,2}(C^*)$ and $\tau_1\tau_2(C^*)$, respectively contain a maximal number of
points in $\overline\alpha_k(0)$, so they are both contained in
$\overline\alpha_k(0)$.
Therefore, we have the following inclusions:
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C^*)\cap M)\subset
\tau_1^{-1}(\tau_1(C^*)\cap M)\cap\tau_2^{-1}(\tau_2(C^*)\cap
M). \subset C^*\cap M\enspace.
\end{equation}
This chain is correct especially if
$\tau_i(C^*)\setminus\overline\alpha_k(0)\neq\emptyset$, $i=1,2$ and
both set are $\alpha_k$-connected.
In both cases the separation property follows.
\end{Beweis}
\begin{Lemma}\label{doppelpunkt1}
It holds:
\begin{enumerate}
\item The set $\alpha_{n-1}(0)\subset\mathbb{Z}^n$ contains no
$\alpha_k$-double points for $0\le k\le n-1$.
\item The set $\alpha_{k}(0)\subset\mathbb{Z}^n$ contains no
$\alpha_{n-1}$-double points for $0\le k\le n-2$.
\end{enumerate}
\end{Lemma}
\begin{Beweis}
1. Let $p$ be in $\alpha_{n-1}(0)$. Then, the point $p$ has the form
$(0,\ldots,0,\pm1,0,\ldots,0)$. It cannot contain any $\pi$-neighbors
$r=(r_1,\ldots,r_n)$ in $\alpha_{n-1}(0)$, because these satisfy
\begin{equation}
\sum_{i=1}^n |p_i-r_i|=1\enspace.
\end{equation}
The point $r$ cannot be 0 and satisfies:
\begin{equation}
\sum_{i=1}^n|r_i|=2\enspace.
\end{equation}
Therefore, no neighbor of $p$ can be contained in $\alpha_{n-1}(0)$ and
no $p$ exists, which satisfies the definition \ref{doppelpunkt}.
2. We need to show, that for no $p\in\alpha_k(0)$ with $0\le k\le n-2$, exist two
points $r\in\pi(0)\cap\alpha_{n-2}(p)$ and $q\in\pi(0)\cap\alpha_k(0)$ and a translation
$\sigma$ with $\sigma(r)=0$ and $\sigma(q)=p$ such that $r\in\alpha_{n-1}(p)$.
Assume for contradiction that such a configuration exists. Then, the two points
$q$ and $r$ are $\alpha_{n-1}$-adjacent. Therefore, it holds:
\begin{equation}
\sum_{i=1}^n|r_i-q_i|=1, |r_i-q_i|\le 1\mbox{ for } 1\le i\le
n\enspace.
\end{equation}
It follows the existence of a $j$ in $\{1,\ldots,n\}$ such that $r_j\neq
q_j$ and $r_i=q_i$ for all other indices $i$. Furthermore, the point $q$ is in $\pi(0)$
and it can be written as $(0,\ldots,\pm1,0,\ldots,0)$ with $q_l=\pm1$ and we know
that $q_l=r_l$. From $\sigma(r)=0$ it follows that $(-\sigma)(q)=p=(q_1-r_1,\ldots,q_n-r_n)$
and therefore, the point $p$ has the form
\begin{equation}
p=(0,\ldots,0,q_j-r_j,0,\ldots,0)\enspace.
\end{equation}
In addition, $r$ is an element of $\pi(p)$ and $|p_i-r_i|\le 1$ for all $1\le i\le n$.
But this cannot be the case, since
\begin{equation}
|p_j-r_j|=|q_l-2r_l|=|-2r_l|=2\mbox{ since } r_l\neq 0\enspace.
\end{equation}
This contradicts the assumption and the Lemma is proven.
\end{Beweis}
\begin{Lemma}\label{doppelpunkt2}
Given a pair $(\alpha_l\alpha_k)$ on $\mathbb{Z}^n$ with $n\ge 2$, the set
$\alpha_k(0)\subset\mathbb{Z}^n$ contains $\alpha_l$-double points for all $0\le
k\le n-2$ and $0\le l \le n-2$.
\end{Lemma}
\begin{Beweis}
Consider the point $p=(1,1,0,\ldots,0)\in\alpha_k(0)$ with $k$ conforming
the precondition.
The point $q=(1,0,\ldots,0)$ is in $\pi(0)$ and the point
$r=(0,1,0,\ldots,0)$ is in $\pi(p)$.
In addition a translation $\sigma$ exists such that $\sigma(z)=r$ and $\sigma(p)=q$.
Because of $q\in\alpha_l(r)$ for $0\le l\le n-2$, the Lemma is true.
\end{Beweis}
We now have all the tools in our hands to state the final Theorem on the good pairs
of cubical ajacencies. This Theorem gives us a complete characterization of this kind of
good pairs in $\mathbb{Z}^n$ for all dimensions $n$ at least 2.
\begin{Satz}
A pair of cubical adjacencies $(\alpha_l,\alpha_k)$ in $\mathbb{Z}^n$ is a good pair, if
\begin{enumerate}
\item $k=n-1$ and $0\le l\le n-2$,
\item $0\le k\le n-2$ and $l=n-1$.
\end{enumerate}
There are no other good pairs of cubical adjacencies.
\end{Satz}
\begin{Beweis}
We need to show that $\alpha_k(0)$ is a $(n-1)$-manifold in $\mathbb{Z}^n$ that contains
no $\alpha_l$-double points. Lemma \ref{doppelpunkt1} gives the pairs
$(\alpha_l,\alpha_k)$ without double points.
Corollary \ref{kub-surf} shows that $\alpha_k(0)$ is a $(n-1)$-manifold under $\alpha_l$,
this is enough because of the invariance under translation of $\alpha_k$.
And from Lemma \ref{doppelpunkt2} we know which pairs of cubical adjacencies have
double points.
\end{Beweis}
\subsection{The Khalimsky-Topology as Good Pair of Adjacencies}\label{section-khalimsky-top}
In this section we will show, that the notion of an Alexandrov-space
and the graph-theoretic framework common to digital geometry may be put under a common
umbrella. We will see, that the Khalimsky-topology $\kappa_n$ on the set $\mathbb{Z}^n$
might be considered as a pair of adjacencies $(\kappa_n,\kappa_n)$, and that these
pairs a good ones.
Basing on Theorem \ref{alexandrov-partial-order} we may consider a graph structure
on $\mathbb{Z}^n$ given by the topology $\kappa_n$. We denote this graphical adjacency also with
$\kappa_n$. Also, remember the equations \ref{eq:closure1} and \ref{eq:closure2}.
\begin{Lemma}
For any $p,q\in\mathbb{Z}^n$ holds:
\begin{equation}
p\in\mbox{\textnormal{C}}_{\kappa_n}(q)\Leftrightarrow\bigwedge_{i=1}^n(p_i\ge q_i\mod{2}) :\Leftrightarrow p\succeq q
\end{equation}
\begin{equation}
p\in\mbox{\textnormal{U}}_{\kappa_n}(q)\Leftrightarrow\bigwedge_{i=1}^n(p_i\le q_i\mod{2}) :\Leftrightarrow p\preceq q
\end{equation}
\end{Lemma}
\begin{Beweis}
This is Theorem 8 in Evako et al. \cite{evako}.
\end{Beweis}
The Khalimsky-adjacency $\kappa_n$ may now be represented in the following way:
\begin{equation}
\kappa_n(p):=\left\{q\in\mathbb{Z}^n: \max_{i=1,\ldots,n}|q_i-p_i|=1,
p\preceq q\lor q\preceq p \right\}\enspace.
\end{equation}
\begin{Lemma}
For all $n\ge 1$ holds: $\pi\subset\kappa_n$.
\end{Lemma}
\begin{Beweis}
Let $p,q$ be two points in $\mathbb{Z}^n$ such that $p\in\pi(q)$. By definition of $\pi$
we have
\begin{equation}
\max_{i=1,\ldots,n}|p_i-q_i|=1\mbox{ and }\sum_{i=1}^n|p_i-q_i|\le 1
\end{equation}
Therefore, exactly one $i\in\{0,\ldots,n\}$ exists with $q_i=p_i+1$ or
$q_i=p_i-1$. For all $j\in\{1,\ldots,n\}$, $j\neq i$ is $p_i=q_i$.
We have:
\begin{equation}
\bigwedge_{i=1}^n(p_i\le q_i\mod{2})\mbox{ or }
\bigwedge_{i=1}^n(p_i\ge q_i\mod{2})\enspace.
\end{equation}
And so, $p\in\kappa_n(q)$.
\end{Beweis}
We are not in the convenient position to find a reference point like 0 for the cubical
adjacencies. The next Lemma clarifies this fact.
\begin{Lemma}
For each $p\in\mathbb{Z}^n$ exists a translation $\tau$ such that
$\kappa_n(\tau(p))\neq \tau(\kappa_n(p))$.
\end{Lemma}
\begin{Beweis}
By construction of the Khalimsky-topology this Lemma is obviously true:
Let $\tau$ be any translation of the form $(0,\ldots,0,1,0,\ldots,0)$ in $\mathbb{Z}^n$.
Then, $\tau(p)$ is either odd in a component where $p$ is even or vice versa.
In both cases, the point $\tau(p)$ has a neighborhood different from the one
of $p$.
\end{Beweis}
We are able to make some statements about the interaction of certain translations and
$\kappa_n$.
\begin{Lemma}\label{kappa-translation}
Let $p$ and $q$ be two points in $\mathbb{Z}^n$, $I=\{i: p_i=q_i\}$ and let
$\tau$ be a translation with$|\tau(0)_i|\le 1$ for $i\in I$ and
$\tau(0)_i=0$ otherwise. Then holds
\begin{equation}
p\preceq q \Leftrightarrow \tau(p)\preceq\tau(q)\enspace.
\end{equation}
\end{Lemma}
\begin{Beweis}
($\Rightarrow$) Let $p\preceq q$. Then holds $p_j\le q_j\mod{2}$ for all
$j\not\in I$. Because of $p_i=q_i$ we deduce $p_i\pm 1= q_i\pm 1\mod{2}$. Therefore,
it holds $\tau(p)\preceq\tau(q)$.
($\Leftarrow$) Analog.
\end{Beweis}
\begin{Lemma}
For all $p\in\mathbb{Z}^n$, $n\ge 2$, the set $\kappa_n(p)$ is
$\kappa_n$-connected.
\end{Lemma}
\begin{Beweis}
The Lemma follows by Definition 4 and Theorem 11 in Evako
et al. \cite{evako}.
\end{Beweis}
From the proof of Theorem 11 in Evako et al. \cite{evako} we get
\begin{Lemma}\label{k:cube-connected}
For all $p\in\mathbb{Z}^n$, $n\ge 2$, every $n$-cube, that contains points from
$\kappa_n(p)$, is $\kappa_n$-connected.\hspace{\stretch{1}}$\Box$
\end{Lemma}
\begin{Lemma}\label{k:two-components}
For all $p\in\mathbb{Z}^n$, $n\ge 2$, and all $q\in\kappa_n(p)$ the set
$\omega(q)\setminus\kappa_n(p)$ has exactly two $\kappa_n$-components
$C_q$ and $D_q$.
\end{Lemma}
\begin{Beweis}
Let $p$ and $q$ be the same as in the last Lemma.
A $\kappa_n$-component of $\omega(q)\setminus\kappa_n(p)$ is $\{p\}$, because $p$
has in $\omega(q)$ only neighbors $\kappa_n(p)$.
We denote this component by $C_q$.
Now define
\begin{equation}
\omega(q)_i := \left\{r\in\omega(q):|r_i-p_i|=2 \right\}
\end{equation}
We will show that this set is $\pi$-connected for all $1\le i\le n$
We prove the result w.l.o.g. for $i=1$.
The point
\begin{equation}
r=(p_1+2,p_2,\ldots,p_n)
\end{equation}
is in $\omega(q)_1$.
Let $r'\neq r$ be any point in $\omega(q)_i$ and let $i\in\{2,\ldots,n\}$
be the smallest index such that $r_i\neq r'_i$.
We construct a $\pi$-path from $r'$ to $r$.
The point
\begin{equation}
r''=(r_1,\ldots,r_i,r'_{i+1},\ldots,r'n)
\end{equation}
is a $\pi$-neighbor of $r'$, because, both points differ according to the choice
of $i$ only in the $i$-th coordinate by 1. If $r''=r$ the the path is constructed,
otherwise we iterate the algorithm with $r''$ in place of $r'$. After at most $n-1$
steps the $\pi$-path is constructed.
If the intersection of two sets $\omega(q)_i$ and $\omega(q)_j$ is non-empty, then
it is $\pi$-connected, too.
Let $r$ be a point in the set
\begin{equation}
D_q:=\omega(q)\cap(\omega((p)\setminus(\kappa_n(p)\cup\{p\}))\enspace.
\end{equation}
If $\omega(q)_i\neq\emptyset$ and $r_i=q_i$ for this
$i\in\{1,\ldots,n\}$, then $r_i$ is the $\pi$-neighbor of some point in $\omega(q)_i$.
Otherwise, an $i$ exists such that $\omega(q)_i\neq\emptyset$. From
$r_i\neq q_i$ follows $|r_i-q_i|=1$ and therefore $r_i=p_i$.
We may define the point
\begin{equation}
s=(r_1,\ldots,r_{i-1},q_i,r_{i+1},\ldots,r_n)\enspace.
\end{equation}
The points $r,s$ are in $\omega(p)$, so we have
\begin{equation}
\max_{i=1,\ldots,n}|r_i-p_i|=1\mbox{ and
}\max_{i=1,\ldots,n}|s_i-p_i|=1\enspace.
\end{equation}
Since the point $r$ is no member of $\kappa_n(p)$, it follows:
\begin{equation}
\bigvee_{j_1}(r_{j_1}>p_{j_1})\mbox{ and } \bigvee_{j_2}(
r_{j_2}<p_{j_2})\enspace.
\end{equation}
The indices $j_1$ and $j_2$ are distinct. From $r_i=p_i$
follows, that $j_1,j_2$ are both dissimilar to $i$.
Therefore we have for $s$:
\begin{equation}
s_{j_1}=r_{j_1}>p_{j_1}\mbox{ and } s_{j_2}=r_{j_2}<p_{j_2}\enspace,
\end{equation}
which gives $s\not\in\kappa_n(p)$. Thereby,
$s$ is in $D_q$ and $D_q$ is the second $\kappa_n$-component of the set
$\omega(q)\setminus\kappa_n(p)$.
\end{Beweis}
\begin{Lemma}\label{k:background-adjacency}
For all $p\in\mathbb{Z}^n$ with $n\ge 2$ and any $q\in\kappa_n(p)$, all the points
$r\in\kappa_n(p)\cap\kappa_n(q)$ are $\kappa_n$-adjacent to the sets $C_q$ and
$D_q$ .
\end{Lemma}
\begin{Beweis}
It obvious, that all points $r\in\kappa_n(p)\cap\kappa_n(q)$ are
$\kappa_n$-adjacent to the set $C_q=\{p\}$. So it remains to show, that $r$ is also
$\kappa_n$-adjacent to $D_q$.
Case 1: For some index $i$ in $\{0,\ldots,n\}$ holds that $r_i=q_i$ and the set
$\omega(q)_i$ is not empty. Then, the point $r$ is $\pi$-adjacent to $D_q$.
Case 2: It is $r_i\neq q_i$ for all $i$ such that $\omega(q)_i\neq\emptyset$.
Consider the set $I=\{i:\omega(q)_i\neq\emptyset\}$. We show that the point
\begin{equation}
s\mbox{ with }
s_i=\left\{
\begin{array}{rcl}
q_i&\mbox{if}& i\in I \\
r_i&\mbox{if}& i\in\{1,\ldots,n\}\setminus I
\end{array}
\right.
\end{equation}
is no member of $\kappa_n(p)$ under this preconditions. Since $s$ can be identified as
$\kappa_n$-adjacent to $r$, the point $r$ is $\kappa_n$-adjacent to $D_p$.
The point $s$ is distinct from $q$ by definition of $I$ and $r\neq p$. We know that $r_i=p_i$
for $i\in I$ because of $r_i\neq q_i$ and $|r_i-p_i|<2$ and $r$ is a member of $\omega(q)$.
The set $\omega(q)_i$ is non-empty if and only if $q_i\neq p_i$. Therefore, a translation
$\tau$ exists such that
\begin{equation}
\tau(0)_i=\left\{
\begin{array}{rcl}
\pm1&\mbox{if}&i\in\{1,\ldots,n\}\setminus I\\
0&\mbox{if}&i\in I
\end{array}\right.
\end{equation}
We may choose $\tau$ such that $\tau(q)=p$ and $\tau(s)=r$.
The point $q$ is in $\kappa_n(p)$. Suppose w.l.o.g. that $q\preceq
p$. Therefore, we get $q_i< p_i=r_i\mod{2}$ for $i\in I$.
Then follows $q\preceq r$ with $r\in\kappa_n(q)$. By definition of
$s$ and the fact $r\neq p$, it holds that $q\preceq s$ and by Lemma
\ref{kappa-translation} we get
\begin{equation}
\tau(q)=p\preceq r=\tau(s)\enspace.
\end{equation}
So we can find a $j$ with $s_j=r_j>p_j\mod{2}$ and $j\not\in I$. But at the same time
$q_i=s_i<p_i\mod{2}$ for all $i\in I$. Therefore the point $s$ cannot be contained in
$\kappa_n(p)$. We have to show that $r$ and $s$ are $\kappa_n$-neighbors: For $i\not\in I$
we have $r_i=s_i$ and for $i\in I$ it holds
\begin{equation}
s_i = q_i < p_i = r_i \mod 2
\end{equation}
and so follows $s\preceq r$ which means $s\in\kappa_n(r)$.
\end{Beweis}
For the proof of the separation property we consider the set
$\overline\kappa_n(p)=\kappa_n(p)\cup\{p\}$ for all $p\in\mathbb{Z}^n$, $n\ge 2$
instead of $\kappa_n(p)$. This is reasonable, since the point $p$ lies in no
separable component of the complement of $\kappa_n(p)$ in $\mathbb{Z}^n$. If we have the
result for the modified set we may easily translate it for the original one.
\begin{Lemma}
Let $C$ be any $k$-cube $\omega(p)\cup\{p\}$, $0\le k\le n$ and let $q$ be the point in
$C$ with minimal $\pi$-distance\footnote{The $\pi$-distance of two points $p$ and $q$ is
the infimum over the length of all $\pi$-paths from $p$ to $q$.} to $p$.
For $q\preceq p$ and all $q'\in C\setminus\{q\}$ holds
\begin{equation}
q'\preceq q\Leftrightarrow q'\preceq p\enspace.
\end{equation}
An analog claim holds for $q\succeq p$.
\end{Lemma}
\begin{Beweis}
($\Rightarrow$) This direction of the proof follows by transitivity of the partial order $\preceq$.
($\Leftarrow$) The points $q'$ and $q$ are contained in the same $k$-cube $C$ and it holds that
\begin{equation}
\sum_{i=1}^n|q_i-p_i|=k<l=\sum_{i=1}^n|q'_i-p_i|\enspace.
\end{equation}
After rearranging the coordinates of $q,q'$ and $p$, we get
\begin{equation}
q'=(q_1,\ldots,q_{k},q'_{k+1},\ldots,q'_l,p_{l+1},\ldots,p_n)
\end{equation}
\begin{equation}
q=(q_1,\ldots,q_{k},p_{k+1},\ldots,p_l,p_{l+1},\ldots,p_n)\enspace.
\end{equation}
From $q'\preceq p$ now follows $q\preceq p$ by the definition of
$\preceq$.
\end{Beweis}
\begin{Lemma}
Let $C$ be a $k$-cube $2\le k\le n$ and $q$ be a point in $C$ with minimal $\pi$-distance to
$p$ and $q\preceq p$. For all $q'\in\pi(q)\cap C$ holds $q'\preceq q$ if and only if $C$ is
contained in the set $\overline\kappa_n(p)$.
An analog claim holds for $q\succeq p$.
\end{Lemma}
\begin{Beweis}
($\Leftarrow$) Since $C\subset\overline\kappa_n(p)$, all the points $q'\in C\cap\pi(q)$ are
in $\kappa_n(p)$. Therefore, they satisfy $q'\preceq q$ or $q'\succeq q$.
If there exists a $q'\preceq q$ and a $q''\succeq q$, so we have
\begin{equation}
\bigvee_{i_1}(q'_{i_1}<q_{i_1}\mod{2})\mbox{ and
}\bigvee_{i_2}(q''_{i_2}>q_{i_2}\mod{2})\enspace.
\end{equation}
The point $q'''=\tau_{i_1}\tau_{i_2}(q)$ then satisfies
\begin{equation}
\bigvee_{i_1}(q'''_{i_1}<q_{i_1}\mod{2})\mbox{ and
}\bigvee_{i_2}(q'''_{i_2}>q_{i_2}\mod{2})\enspace.
\end{equation}
Therefore holds $q'''\not\preceq q$ and $q'''\not\succeq q$ and the point $q'''$
no member of $\overline\kappa_n(p)$. So for all points $q'\in C\cap\pi(q)$
the relation $q'\preceq q$ holds.
($\Rightarrow$) We prove by induction on $k$.
In the case $k=2$ holds $q\preceq p$ and for all $q'\in C\cap\pi(q)$ holds
$q'\preceq q\preceq p$. The two $\pi$-neighbors $q_1$ and $q_2$ of $q$ in $C$ are in
$\overline\kappa_n(p)$. This means that
$q_1=\tau_1(q)\preceq q$ and $q_2=\tau_2(q)\preceq q$.
Therefore, we have
\begin{equation}
\tau_1(\tau_2(q))\preceq \tau_1(q)\preceq p\enspace.
\end{equation}
We conclude, that $C$ is contained in $\overline\kappa_n(p)$.
For the induction step $k>2$ we let $C=C'\cup\tau(C')$ for certain $(k-1)$-cubes $C',\tau(C')$
and a translation $\tau$. Let $q$ be in $C$ w.l.o.g.
Since all the points $q'$ in $\pi(q)\cap C$ satisfy the relation
$q'\preceq q$, the $(k-1)$-cube $C'$ has to be contained by induction hypothesis in
$\overline\kappa_n(p)$. For all $q''\in C'$ holds $q''\preceq q$.
Therefore, by Lemma \ref{kappa-translation}, we find for all $\tau(q'')\in\tau(C')$:
\begin{equation}
\tau(q'')\preceq\tau(q)\preceq q\preceq p\enspace.
\end{equation}
It follows that $C\subset\overline\kappa_n(p)$.
\end{Beweis}
\begin{Folg}
Let $C$ be a $k$-cube, $2\le k\le n$ and $q$ be the point with
minimal $\pi$-distance to $p$. Then, all the subcubes $C'$ of $C$ such that
$q'\preceq q\preceq p$ or $q'\succeq q\succeq p$ for all $q'\in
C'\cap\pi(q)$, are contained in $\overline\kappa_n(p)$.
For $q\neq p$ only one of these cases applies.\hspace{\stretch{1}}$\Box$
\end{Folg}
\begin{Lemma}\label{kappa-trenn1}
The set $\overline\kappa_n(p)$ has the separation property under the pair
$(\kappa_n,\kappa_n)$ for any cube $C\subset(\omega(p)\cup\{p\})$ with $p\not\in C$.
\end{Lemma}
\begin{Beweis}
We consider three cases. The first case is, that $C$ is contained in
$\overline\kappa_n(p)$. The separation property is obviously satisfied in this case.
Case 2: Let the $k$-cube $C$ be of the form $C'\cup\tau(C')$ with $C'$ a $(k-1)$-cube
contained in $\overline\kappa_n(p)$. In this case the set $\tau(C')$
contains no points $q'$ in $\overline\kappa_n(p)$, since otherwise these points would
satisfy $q'\le q\le p$. Particularly, the point $\tau(q)$ is not in $\overline\kappa_n(p)$.
Every $(k-2)$-cube $C''$ with $C''\cap\overline\kappa_n(p)$ is in $C'$.
Then, the set $\tau_1(C'')\subset C'$ is also contained in
$\overline\kappa_n(p)$. Therefore the separation property holds in $C$.
Case 3: There is only one $(k-2)$-subcube $C'$ of $C$ that contains all points of
$C\cap\overline\kappa_n(p)$. Then we get
$\tau_1\tau_2(C')\cap\overline\kappa_n(p)=\emptyset$. Therefore it holds
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C')\cap\overline\kappa_n(p))
\subset (\tau_1^{-1}(\tau_1(C')\cap\overline\kappa_n(p))) \cap
(\tau_2^{-1}(\tau_2(C')\cap\overline\kappa_n(p)))\enspace.
\end{equation}
And so, the separation property holds.
\end{Beweis}
\begin{Lemma}\label{kappa-trenn2}
The set $\overline\kappa_n(p)$ has the separation property for the pairs
$(\kappa_n,\kappa_n)$ for cubes $C\subset\omega(p)\cup\{p\})$ with $p\in C$.
\end{Lemma}
\begin{Beweis}
Case 1: The separation property is satisfied for $C\subset\overline\kappa_n(p)$.
Case 2: For a $k$-cube $C$ of the form $C'\cup\tau(C')$ such that
$C'\subset\overline\kappa_n(p)$ only the point $\tau(p)$ is in
$\overline\kappa_n(p)$, because, if for all $q\in C'$ the relation
$q\succeq p$ is true, then it holds for $\tau(q)\in\tau(C')$ that
\begin{equation}
\tau(q)\succeq\tau(p)\preceq p\enspace.
\end{equation}
Since $\tau(p)$ has minimal $\pi$-distance to $p$ in $\tau(C')$, none of the aforementioned
$\tau(q)$ can be contained in $\kappa_n(p)$.
Let $C''\subset C'$ be any $(k-2)$-cube. Then, the set $C''\cap\overline\kappa_n(p)$ is
maximal with respect to inclusion in $C$. In turn, the set
$\tau_1(C'')\setminus \overline\kappa_n(p)$ is empty and the separation property holds
for $C$.
Case 3: Consider the $k$-cube $C=C'\cup\tau_1(C')\cup\tau_2(C')\cup\tau_1\tau_2(C')$
and let $C'\cap\overline\kappa_n(p)$ be maximal with respect to inclusion.
Since we are not in case 2, we have
$\tau_1(C')\setminus\overline\kappa_n(p)\neq\emptyset$. The $(k-1)$-cube $C'$
has a $l$-subcube, $0\le l<k-1$, that is contained in $\overline\kappa_n(p)$,
the point $p$ has to be in $C'$ liegen. Now, either all points $q\in C'$ are in
relation $q\preceq p$ or they satisfy $q\succeq p$.
W.l.o.g. we use the first relation.
All the points in $q\in\tau_i(C')$, $i=1,2$, are in the relation $q\succeq p$,
since otherwise, we had $C'\cup\tau_i(C')\subset\overline\kappa_n(p)$.
Now, we have $\tau_1\tau_2(p)\succeq\tau_1(p)\tau_2(p)\succeq p$. Likewise, all the
translations $\tau$, that generate the $(k-1)$-cube $C'$, satisfy by Lemma
\ref{kappa-translation}:
\begin{equation}
\tau_1\tau_2(\tau(p))\succeq \tau_1(\tau(p)),\tau_2(\tau(p))\succeq \tau(p)
\end{equation}
Therefore, only the points $\tau_1\tau_2(\tau(p))$ and $\tau_i(\tau(p))$, $i=1,2$
are in $\overline\kappa_n(p)$, if $\tau(p)\succeq p$ holds. So we have
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C')\cap\overline\kappa_n(p))
\subset (\tau_1^{-1}(\tau_1(C')\cap\overline\kappa_n(p))) \cap
(\tau_2^{-1}(\tau_2(C')\cap\overline\kappa_n(p)))
\end{equation}
and the separation property holds in $C$.
\end{Beweis}
\begin{Folg}\label{k:separation-property}
The set $\overline\kappa_n(p)$ has the separation property under the pair $(\kappa_n,\kappa_n)$.
\end{Folg}
\begin{Beweis}
The claim follows from the Lemmata \ref{kappa-trenn1} and
\ref{kappa-trenn2} for cubes $C\subset\omega(p)\cup\{p\}$.
For any cube $C$ that is not contained in $\omega(p)\cup\{p\}$, the separation property
holds, because $C$ has the form $C'\cup\tau_1(C')\cup\tau_2(C')\cup\tau_1\tau_2(C')$
and the set $\tau_1\tau_2(C')\cap\overline\kappa_n(p)$ is always empty, since $C'\subset
\omega(p)\cup\{p\}$ is true if we maximize the set $C'\cap\overline\kappa_n(p)$ with
respect to inclusion. In the case
$\tau_1(C)\setminus\overline\kappa_n(p)=\emptyset$ the separation property holds trivially.
For $\tau_1(C)\setminus\overline\kappa_n(p)\neq\emptyset$ this is also true because of
\begin{equation}
(\tau_1\tau_2)^{-1}(\tau_1\tau_2(C')\cap\overline\kappa_n(p))=\emptyset
\subset (\tau_1^{-1}(\tau_1(C')\cap\overline\kappa_n(p))) \cap
(\tau_2^{-1}(\tau_2(C')\cap\overline\kappa_n(p)))\enspace.
\end{equation}
And so the separation property holds again.
\end{Beweis}
\begin{Satz}\label{kappa-mannigfaltigkeit}
For all $p\in\mathbb{Z}^n$, $n\ge 2$ the set $\kappa_n(p)$ is a
$(n-1)$-manifold.
\end{Satz}
\begin{Beweis}
The first three properties of a digital $(n-1)$-manifold
are shown in the Lemmata \ref{k:cube-connected} to \ref{k:background-adjacency} and the
separation property is proven in Corollary \ref{k:separation-property}.
\end{Beweis}
\begin{Lemma}\label{kappa-doppel}
Given the pair $(\kappa_n,\kappa_n)$ on $\mathbb{Z}^n$, $n\ge 2$, and any point $p\in\mathbb{Z}^n$, the set
$\kappa_n(p)$ contains no $\kappa_n$-double points.
\end{Lemma}
\begin{Beweis}
Assume for contradiction, we have the points $z\in\kappa_n(p)$, $q\in\kappa_n(p)\cap\pi(z)$ and
$r\in\kappa_n(z)\cap\pi(p)$, and $q=\sigma(p)$ and $z=\sigma(r)$ for a simple
translation $\sigma$.
The point $z$ is in $\kappa_n(p)$ and so we have $z\preceq p$ or $p\preceq
z$. We consider w.l.o.g. the case $z\preceq p$.
We have exactly one $i\in\{1,\ldots,n\}$ such that
\begin{equation}
z_i=r_i<p_i\mod{2}\enspace.
\end{equation}
Therefore, it holds that
\begin{equation}
q_i=p_i>r_i\mod{2}\enspace.
\end{equation}
Furthermore, we can find a $j\in\{1,\ldots,n\}$ such that
\begin{equation}
z_j=q_j<p_j=r_j\mod{2}\enspace.
\end{equation}
It follows that $q_i>r_i\mod{2}$ and $q_j<r_j\mod{2}$. Therefore neither
$q\preceq r$ nor $r\preceq q$ may be true. This contradicts the assumption that
$q\in\kappa(p)$ and so no double points may occur.
\end{Beweis}
\begin{Satz}
The pair $(\kappa_n,\kappa_n)$ is a good pair on $\mathbb{Z}^n$ for
all $n\ge 2$.
\end{Satz}
\begin{Beweis}
The proof follows with Theorem \ref{kappa-mannigfaltigkeit} and
Lemma \ref{kappa-doppel}.
\end{Beweis}
\section{Conclusions}
We have shown that the cubical adjacencies and the khalimsky-topology give good pairs.
This was already known, for instance G.T.~Herman proved this in his book \cite{herman}.
The difference here is, that our theory resembles more closely the euclidean case and surfaces
are really subsets of the given space. We also could give a slight unification of the
topological with the graph-theoretic setting, although this was already present in the
disguise of Alexandrov-spaces, for these have an graph-theoretic interpretation via
partial orders. It is possible to give proofs for other adjacency relations to be good
pairs, for instance the hexagonal adjacencies also give good pairs, as G.T. Herman
shows in the same book. It may be also possible to give good pairs of more complicated
adjacency relations, but then, the proofs might tend to get even more technical than
the ones we saw we saw in this paper.
|
\section{Introduction}
The center of the Milky Way galaxy (MW) is an extreme environment, very different from the other locations in the MW. It contains a dormant $\sim 4\times 10^6 M_{\odot}$ black hole, three massive clusters of young high-mass stars and a giant molecular cloud (see reviews in \cite{schdel etal06}). The densities of gas and dust are 2-3 orders of magnitudes higher than the Galactic average, the magnetic field is in the range of
$(100\div 1000)\mu G$, while the Galactic norm is few $\mu G$.
The gas exhibits a complex filamentary morphology as well as a supersonic velocity field.
\begin{figure}[h]
$
\begin{array}{cc}
\includegraphics[scale=0.8]{Spitzer.eps} &
\includegraphics[scale=0.4]{sig11-003_Inline.eps}
\end{array}$
\caption{{\bf left}: Spitzer Telescope--artist concept. {\bf right}: Spitzer image of the Galactic center region. {\it blue}: 3.6 $\mu$ m ,{\it green}: 8 $\mu$ m ,{\it red}: 24 $\mu$ m . The brightest white feature at the center of the image is the central star cluster in our galaxy, surrounding the massive black hole. {\it source}: Spitzer website.}
\label{GC}
\end{figure}
The GC has been observed in infrared and radio bands because the high gas and dust densities combined with a line of sight to the Galactic center (GC) that passes through the MW disk, strongly absorb optical and UV radiation. Infrared telescopes have advanced considerably in the past decades; the Spitzer infrared space telescope is the most recent and advanced infrared facility to date. The left panel of
Figure \ref{GC} is an artist concept of the Spitzer telescope while the right one is the Spitzer image (in 3 mid-infrared wavelengths) of a $\sim 600 $ pc extent of the MW disk around the GC.
\begin{figure}[h]
\centerline{\includegraphics[scale=0.6]{contini09_f1.eps} }
\caption{ A 21 cm radio continuum (log-log scale) image \cite{yusef_morris87} with 11 arcsec resolution.Except for the Radio Arc and part of the Sgr A region which are non-thermal, all the radio emission is thermal bremsstrahlung. The 38 positions observed by with the Spitzer telescope in the mid- infrared \cite{simps07}are indicated by square boxes.(adapted from Fig. 1 of \cite{simps07}) }
\label{simps}
\end{figure}
Zooming in, Figure \ref{simps} which is adapted from \cite{simps07}, shows the inner 50 pc of the GC. The black hole Sgr A West is surrounded by the nuclear star cluster (not marked in the figure). At a distance of about 25 pc in the Galactic plane are the Quintet and the Arches star clusters. Most of the stars are young (age $ \sim 2$ Myr) and massive (10 - 50)~$M_{\odot}$. Visible in the images are the Arches and the Sickle gas filaments. Visible also is a system of quasi circular arcs whose morphologies suggest that they were formed by stellar winds (velocities $\sim 1000$~km/s) emanating from the Quintet and the Arches star clusters.
Observations with the Spitzer Telescope
by \cite{simps07} yielded high quality mid-infrared spectra at 38 positions along a line perpendicular to the MW disk (the positions are marked by the squares in Figure \ref{simps}). The observational spectra lead to the conclusion that the gas in the filaments is indeed excited by ionizing UV photons
originating in the adjacent star clusters \cite{simps07}. However, photo-ionization by itself cannot account for the observed line ratios and an additional excitation by shocks is probably important.
Shocks are indeed expected since
the line of sight (radial) gas velocities measured by radio observations (\cite{yusef_etal97, lang_etal05}) span a range of 180~km/s. Moreover, the filamentary fractured structure of the gas can be naturally explained as the result of turbulence generated by shocks.
Following on this proposition, a detailed modeling
of the line and continuum spectra allowing for both photo-ionization and shocks as the sources of excitation has been carried out \cite{contini09} . All the available lines in each spectrum and the continuum spectral energy distribution were modeled consistently.
Indeed, this modeling reproduced correctly the observational spectra and at the same time was able to derive the shock velocity, pre-shock density and magnetic field, and the geometrical depth at each of the 38 positions. More recently \cite{contini_goldman11},
these values of the physical parameters have been used
to calculate the expected spectra in the optical and UV ranges, which cannot be observed by the presently available instruments.
\section{Outline}
The analysis presented in \cite{contini09} invoked the existence of shocks as a necessary ingredient
for the consistent derivation of the spectra observed by \cite{simps07}.
Shocks are known to produce turbulence, notably via the Richtmyer-Meshkov instability (\cite{mikael90}, \cite{graham_zhang})
which is due to the shock acceleration, and
is similar to the familiar Raleigh-Taylor instability. Shocks can also excite turbulence by creating flows that exhibit
shear instability or Kelvin Helmholtz instability which in turn generate turbulence.
Indeed, the clumped and fragmented morphology of the region strongly suggests the existence of an underlying supersonic shock-generated turbulence.
The existence of turbulence can be probed directly by analyzing the velocity spectrum. In addition, the imprint of turbulence may be detected indirectly by analyzing the power spectrum of
"passive scalars"(\cite{lesieurs97}), which are
strongly coupled to the gas and follow its turbulent motion (but do not feed-back on the
turbulence). The power spectrum of a passive scalar is
proportional to that of the turbulent velocity.
An example is the 21 cm emissivity which is proportional to the column density of the neutral hydrogen and its fluctuations reflect the density fluctuations and in turn the velocity fluctuations. In the case of infrared continuum the passive scalar is the dust density, and in the case of abundances it is the concentration of the species under consideration.
The observational power spectrum of a passive scalar provides a measure of its
hierarchical spatial structure. When this power spectrum is a power law, a hydro-turbulence is naturally suggested as the
mechanism that has shaped the observed spatial hierarchy. Yet, by itself, such a spectrum is not enough to prove the existence of a hydrodynamical turbulence. On the other hand, if the power spectrum of the velocity field reveals a turbulence, the fact that
the power spectrum of a passive scalar is also a power law with the same exponent strengthens the case
for the reality of the velocity turbulence.
An example is the power spectrum of
21cm emission in the Small Magellanic Cloud (SMC) \cite{stanimir99}. It was shown by \cite{goldman2000} that the power spectrum is consistent
with that generated by a large scale velocity turbulence.
Additional support for the existence of a dynamic turbulence was obtained by \cite{goldman07} by deriving the power spectrum of
the radial velocities of the giant H$_ I $ super-shells of the SMC and showing that the two power spectra are consistent.
In what follows we analyze the radial velocity data of \cite{simps07} and their observed
mid-IR continuum flux in order to test for the existence of turbulence.
Indeed these two kinds of observational data reveal the existence of a supersonic turbulence.
Following these two analyses, we examine the Si/H abundance ratio as function of position which were {\it computed} by \cite{contini09} . The idea being that if these computations indeed reproduce the actual physical conditions in the region, then
the imprint of the turbulence should be evident in the computed abundance as function of position.
It turns out to be indeed the case; thus lending
credibility to
the computations of \cite{contini09} as well as to, their extensions to the optical and UV ranges \cite{contini_goldman11}.
Finally, we estimate the effect of turbulence on the magnetic field in the observed positions.
\section{Turbulence: Radial Velocities}
The 3D spectral function of the turbulent velocity, $\Phi(\vec{k})$ is defined in terms of the 2-point
autocorrelation of the turbulent velocity field
$$<\vec{v}(\vec{r'}) \cdot \vec{v}(\vec{r}+\vec{r'})>= \int
\Phi(\vec{k})e^{i \vec{k}\cdot \vec{r}} d^3k$$
In the homogeneous and isotropic case it is useful to introduce the
turbulence energy spectrum $E(k)$ and the turbulent velocity spectral function $F(k) = 2 E(k)$ so that
$$\Phi(\vec{k})=\Phi(k)=\frac{ F(k)}{4\pi k^2}\ ; k=| \vec k|$$
Assuming the ergodic principle, ensemble averages can be replaced by space, surface, or as in the present case, line averages.
The assumed isotropy implies that $F_r(k)$, - the power spectrum of the radial turbulent velocity is
$$F_r(k) =\frac{1}{3} F(k)$$
and can be derived from
$$<v_r(x')v_r(x'+x)>= \frac{1}{L}\int_0^L v_r(x')v_r(x'+x)dx = \frac{1}{2\pi}\int_0^L F_r(k)e^{ik x} dk$$
where $L$ is the length of the line along which the radial velocity has been measured.
The power spectrum thus is the one dimensional Fourier transform of the 2-point
autocorrelation function of the radial velocity. Since the data are given at a set of discrete positions we compute the power spectrum by employing a digital Fourier transform.
Note that at each position the mean squared turbulent radial velocity
$$ <v{_r}^2> = \frac{1}{2\pi}\int_0^L F_r(k) dk$$
is contributed by all the spatial scales of the spectrum.
The observational data are given for 38 positions along an almost straight line
with extension of about 75~pc \cite{simps07}. Not all positions are evenly spaced, thus a numerical
uncertainty in the small-scales part of the power spectrum is expected.
\begin{figure}[h]
\centerline{\includegraphics[scale=1]{turb1_err.eps}}
\caption{The power spectrum of the radial velocity residuals, in units of (km/s)$^2$, as function of the relative wavenumber k. $k=1$ corresponds to the spatial scale $l_0= 75 pc$. The line is a power law with index -2.}
\label{vel}
\end{figure}
Also, the radial velocity observed by \cite{simps07} for the m20 position is substantially larger than the
adjacent velocities and most likely is a foreground flow. We therefore adopt for this point a value equal to the mean velocity of the other positions.
If the residual velocities were random {\it uncorrelated} fluctuation the power spectrum would have been a "white noise" with the same power on all wavelengths.
In contrast,
the power spectrum shown in Figure \ref{vel} is a rather steep decreasing function of the wavenumber $k$.
The largest turbulence scale corresponding to the dimensionless wavenumber $k=1$ is the linear extent of the data strip: $l_0=75$~pc.
The value of the root mean squared turbulent radial velocity at the largest scale is $v_0 = 17.4$~km/s corresponding to a 3D velocity of 30~km/s, well above the
thermal velocity of $\sim 10$~km/s.
The line in the figure is a $k^{-2}$ power law that is expected for supersonic turbulence (\cite{passot88}, \cite{girimaj95}). It differs from
$k^{-5/3}$ Kolmogorov spectrum which characterizes incompressible turbulence. The steeper slope
is due to the fact that a fraction of the turbulent kinetic energy density is converted to compression work and is diverted from the energy cascaded to the smaller spatial scales.
In order to obtain a quantitative estimate of the goodness of the fit we adopted the uncertainty
of 15~km/s for each individual measurement \cite{simps07}. Then we obtained synthetic velocity data
by generating at each position a random velocity using a normal distribution with a mean equal
to the observed value, and a standard deviation of 15~km/s.
We have generated 100 such simulated data sets. For each, the power spectrum was computed enabling an estimate for the standard deviation of the power spectrum.
Doing so, a value of 0.72 was obtained for the reduced $\chi^2$ . Some of the uncertainty in the power spectrum stems from
the uneven spacing of the positions leading to a scatter in the power spectrum at the small spatial scales (large k). The fact that even so the fit is a good one suggests that
the true uncertainty in the observed velocities is {\it smaller} than the assumed value of 15~km/s .
The characteristic timescale of the turbulence is given by $\tau_0\sim l_0/v_0 \sim 4 \times 10^6\ yr$ --
comparable to the age of the young stellar populations.
This is consistent with the framework adopted in \citep{simps07} and \cite{contini09} that
stellar outflows are the main energy source that shaped the interstellar medium in this region and together with the stellar UV radiation are also responsible for the observed mid-infrared spectra .
The derived power spectrum seems to steepen to a $k^{-3}$ dependence at about $k=7$. If real, it implies that geometrical depth along the radial line of sight
is about $\frac{1}{15}$ of the extent on the plane of the sky (\cite{goldman2000}); namely a depth of $\sim 5\ pc$.
\section{Turbulence: Mid-IR Flux}
The observed flux is an integral of the emissions along the line of sight. In the optically thin case
it is contributed by all the geometrical depth along the line of sight while in the optically thick case only the foreground layers contribute. Recently, \cite{mivil07, mivil10}
have studied fluctuations of the far-IR continuum flux from thermally heated dust and concluded that their power spectrum is identical to that of the underlying velocity turbulence. The interpretation of this result was that the dust is strongly coupled to the gas and thus the dust density fluctuations are determined by the gas velocity field which is turbulent. In the cases considered by these authors the turbulence was subsonic and the power spectrum was the Kolmogorov spectrum.
Pursuing a similar analysis, we compute the power spectrum of the fluctuations of the mid-IR continuum flux in the range $(13.5-14.3)\mu m$, as function of position. Mathematically, the procedure is the same employed in the previous section with regard to the radial velocity.
The resulting power spectrum is displayed in Figure \ref{ir}. It exhibits a clear $ k^{-2}$ behavior for the larger scales and an indication of a possible steepening at about $ k=7$ to a $k^{-3}$ dependence; in effect the latter is more obvious here than in the radial velocity spectrum.
The uncertainties reported by \cite{simps07} are
quite small (of the order of $1\%$ ). As a result the uncertainties in the power spectrum
are too small to be indicated in the figure. The scatter in the power spectrum is dominated by the uneven spacing of the positions \cite{simps07}. The latter can be estimated
by generating synthetic data with randomly spacings
(much in the way as with the velocity uncertainties in the previous section),
However this seems to be outside the scope of the present paper.
Therefore, we do not derive here a quantitative estimate for the goodness of fit.
Rather we would like to draw attention to the fact that the deviations from the power law are mostly at the small spatial
scales, as expected from an uneven spacing, and not on the large scales.
The important conclusion from this power spectrum is that it is consistent with the interpretation that the dust density fluctuations (responsible for the flux fluctuations) were driven by the supersonic velocity turbulence addressed in the previous section.
\begin{figure}[h]
\centerline{\includegraphics[scale=1]{turb2.eps}}
\caption{The power spectrum of the mid-IR continuum at $(13.5-14.3)\mu m$, in units of Jy$^2$, as function of the relative wavenumber k. $k=1$ corresponds to the spatial scale $l_0= 75 pc$. The line is a power law with index -2.}
\label{ir}
\end{figure}
\section{Turbulence: Si/H Abundance Ratio}
In \cite{contini09} various abundance ratios were computed as function of position.
If these computations are a fair representation of the physics in the region, it is expected that the computed abundances will reflect the turbulence revealed in the observational radial velocity and mid-IR flux.
If the heavy elements, whose abundance is considered, are
strongly coupled to the gas and follow the gas motions then they can be regarded as a turbulent "passive scalars" (\cite{lesieurs97}), much in the same way as the dust density responsible for the mid-IR emission. As in the latter case, the power spectrum should be
proportional to the turbulent velocity power spectrum.
We computed the power spectrum
of the fluctuations of the Si/H ratio at the 38 positions mentioned in the previous sections. This specific abundance was selected because it does not show drastic variations with position and thus (unlike as with e.g. Fe/H) the fluctuations are not affected by trapping into dust grains and eventual sputtering.
The resulting power spectrum is shown in Figure \ref{si}. The uncertainties in the power spectrum resulting from the uncertainties of the computed abundance are too small to be indicated in the figure. The scatter seen in the small scales (large wavenumbers) probably reflects the uneven spacing of the observational positions.
The power spectrum is consistent with the $k^{-2}$ spectrum of the radial velocity turbulence and that of the mid-IR flux. The consistency of the power spectrum with what is expected from an underlying supersonic hydrodynamic turbulence strengthens
the case for the existence of the latter.
More importantly, this lends credibility to the compuational model employed in \cite{contini09}.
Here, too there is an indication for a steepening of the spectrum at $ k \sim 7$ to a $k^{-3}$ dependence.\\
\begin{figure}[h]
\centerline{\includegraphics[scale=1]{turb3.eps}}
\caption{ The power spectrum of Si/H residuals, in units of $10^{-10}$, as function of the relative wavenumber k. $k=1$ corresponds to the spatial scale $l_0= 75 pc$. The line is a power law with index -2.}
\label{si}
\end{figure}
\section{Discussion}
The clumped and fragmented morphology of the GC region strongly suggests that it has been shaped
by a supersonic turbulence. The latter has been generated by shocks driven by the high velocity stellar winds from the young massive stars located in the adjacent star clusters.
Shocks are known to produce turbulence directly, via the Richtmyer-Meshkov instability
due to the shock acceleration, and also indirectly by giving rise to
shear and Kelvin Helmholtz instabilities which in turn excite turbulence.
The existence of such a turbulence was revealed in the power spectra of the
observational radial velocities and mid-IR continuum flux. The power spectra exhibit a $k^{-2}$ behavior typical to supersonic turbulence, and is steeper
than the Kolmogorov spectrum that corresponds to incompressible turbulence.
A steepening of the power spectra for relative wave numbers exceeding 7, suggests that the line of
sight depth of the turbulent region is about 5pc.
The turbulence radial root mean squared velocity is about 17~km/s and its 3D value is about 30~km/s indicating a turbulence Mach number of about 2-3.
The associated turbulence timescale is about 4 Myr -- comparable to the age of the young
stellar populations which are ultimate generators of the shocks and the ensuing turbulence.
Since the modeling of the observed mid infrared spectra \cite{contini09} invokes the existence of shocks-- our finding
lends credibility to the latter assumption. The models also predict the abundances as function of position \cite{contini09}. Analysis of the Si/H abundance field, reveals the same underlying turbulence as do the observational velocities and the mid-infrared flux. This implies that the modeling of \cite{contini09} does indeed reproduces the true physical
parameters of the gas.
The turbulence can amplify the initial magnetic fields \cite{contini09} up to equipartition value if the available time is much larger than the turbulence timescale. The amplification
proceeds via the dynamo mechanism involving the winding up of magnetic field flux lines by the turbulence eddies.
For the 3D turbulent velocity of 30~km/s , the equipartition value for the largest scale is
$B_{turb,eq}=140 \mu G \ (n/ 100cm^{-3})^{1/2}$
with $n$ denoting the gas number density.
Fields of this strength and even an order of magnitude higher were reported for the Galactic
center region (see review by \cite{Val} ). The larger fields could have evolved from the turbulently
amplified field by an inverse cascade process (\cite{mininni07}) from small to large spatial scales, or directly by shock compression of the
magnetic flux lines embedded in the
ionized ISM \cite{medvedev07}.
\begin{theacknowledgments}
Itzhak Goldman thanks Michael Mond and Pierre-Louis Sulem for the invitation to the interesting and well organized WISAP-2011 conference in Eilat. He also wishes to thank
the Afeka College Research Authority for
financial support.
\end{theacknowledgments}
\bibliographystyle{aipproc}
|
\section{Summary of previous work}
Let $X_1, X_2, \ldots, X_N$ be i.i.d. real valued random variables with distribution function $F$ and corresponding realized values $x_1, x_2, \ldots, x_N$. In the remainder of the paper we assume that $n \in \{1, 2, \ldots, N\}$. The realized value $x_n$ of the random variable $X_n$ is either an exact observation or censored into an interval $(t_n, t_{2n}]$. We allow for the possibility that $t_{2n}=\infty$ and adopt the convention that $(t_n, t_{2n}]$ is to be interpreted as $(t_n, \infty)$ in that special case.
For a given element $\tau \in \mbox{dom}(F)$ we define $d_\tau$ as the number of sample values observed to be less than or equal to $\tau$ and $a_\tau$ as the number of sample values observed to be greater than $\tau$. The count $u_\tau$ represents the number of censored sample values with censoring intervals that capture $\tau$. For example, a censored value $x_n$ is included in the count $d_\tau$ iff $t_{2n}\leq\tau$ and in the count $a_\tau$ iff $t_n\leq\tau$. From these definitions immediately follows that for any $\tau \in \mbox{dom}(F)$ we have that $d_\tau+a_\tau+u_\tau = N$.
Let $k_\tau$ be the actual number of sample values not exceeding $\tau$. Due to the presence of the censoring mechanism the value of $k_\tau$ is only observed to satisfy $d_\tau \leq k_\tau \leq d_\tau+u_\tau$; we label the latter event as $E$. The likelihood of $E$ is given by
\[L(F(\tau); E) = \sum^{d_\tau + u_\tau}_{k_\tau = d_\tau} {N \choose k_\tau}\left[F(\tau)\right]^{k_\tau} \left[1-F(\tau)\right]^{N-k_\tau}\]
Let us define the function $\hat{F}: \mbox{dom}(F) \rightarrow [0, 1]$ as the value of $p$ that maximizes
\[L(p; E) = \sum^{d_\tau+u_\tau}_{k=d_\tau} {N \choose k}(p)^k (1-p)^{N-k}\]
subject to the constraint $0 \leq p \leq 1$. The value of $\hat{F}(\tau)$ has been derived to be
\[ \hat{F}(\tau) = \left\{
\begin{array}{lll}
0 & \mbox{if} & d_{\tau}=0 \mbox{ and } a_{\tau}\geq 1 \\
\\
1 & \mbox{if} & a_{\tau}=0 \mbox{ and } d_{\tau}\geq 1 \\
\\
1/2 & \mbox{if} & u_{\tau}=N \\
\\ \left(1+\sqrt[u_{\tau}+1]{\frac{a_{\tau}(a_{\tau}+1)\ldots(a_{\tau}+u_{\tau})}{d_{\tau}(d_{\tau}+1)\ldots(d_{\tau}+u_{\tau})}}\,\right)^{-1}&\mbox{o.w.}
\end{array}
\right. \]
Furthermore, the function $\hat{F}$ can be used as an estimator for $F$ since it is a non-decreasing function over $\mbox{dom}(F)$.
\section{Multivariate extension}
In this section the definition of the estimator $\hat{F}$ has been extended to the general case of a sample of $M$-variate observations. Let $\bold{X}_1, \bold{X}_2, \ldots, \bold{X}_N$ be i.i.d. $M$-vectors with distribution function $F$ and the matrix $\bold{D}$ be defined as
\[ \bold{D} =
\left( \begin{array}{c}
\bold{X}_1\\
\bold{X}_2\\
\vdots\\
\bold{X}_N
\end{array} \right) =
\left( \begin{array}{cccc}
X_{11} & X_{12} & \ldots & X_{1M}\\
X_{21} & X_{22} & \ldots & X_{2M}\\
\vdots & \vdots & & \vdots\\
X_{N1} & X_{N2} & \ldots & X_{NM}
\end{array} \right)
\]
For the rest of the paper we have assumed that all observations $X_{nm}$ are censored into corresponding intervals $(L_{nm}, R_{nm}]$ since the treatment of a dataset $\bold{D}$ containing exact in addition to censored observations does not provide any new mathematical insight.
We also adopt the convention that unless explicitly stated otherwise, an index represented by a small letter ranges between $1$ and the value of the corresponding capital letter inclusive. For example, $m \in \{1, 2, \ldots, M\}$. Furthermore, a random quantity will be always designated by a capital letter and the corresponding small letter will be reserved for its realization. For example, $\bold{x}_n$ is the realization of the random vector $\bold{X}_n$.
Let $\tilde{X}_{i_m}^{(m)}$ be the value of the $i_m$-th biggest element, $i_m \in \{1, 2, \ldots, I_m\}$, of the set
\[ \left\{ L_{1m},\, L_{1m}, \dots, L_{Nm} \right\} \cup \left\{ R_{1m},\, R_{1m}, \dots, R_{Nm} \right\} \]
and the set $G^{(m)}$ be defined as
\[ G^{(m)} = \left\{ \tilde{X}_1^{(m)},\, \tilde{X}_2^{(m)}, \ldots, \tilde{X}_{I_m}^{(m)} \right\} \]
Consequently, the elements of $G^{(m)}$ are all distinct and such that
\[ \tilde{X}_1^{(m)} < \tilde{X}_2^{(m)} < \ldots < \tilde{X}_{I_m}^{(m)} \]
Let us also define the grid $G$ as $G = G^{(1)} \times G^{(2)} \times \ldots \times G^{(M)}$. Our goal will be to estimate $F$ over $G$.
Let $\bold{x} = (x_1, x_2, \ldots, x_M) \in R^M$ and $\bold{x}' = (x'_1, x'_2, \ldots, x'_M) \in R^M$. We will write $\bold{x} < \bold{x}'$ iff $x_m < x'_m$. The expressions $\bold{x} > \bold{x}'$, $\bold{x} \leq \bold{x}'$ and $\bold{x} \geq \bold{x}'$ are defined analogously. Let $\bold{L}_n = (L_{n1}, L_{n2}, \ldots, L_{nM})$ and $\bold{R}_n = (R_{n1}, R_{n2}, \ldots, R_{nM})$. By analogy with the $1$-dimensional case we define $d(\bold{x})$ as the count of observations $\bold{X}_n$ such that $\bold{R}_n \leq \bold{x}$ and $u(\bold{x})$ as the count of observations satisfying $\bold{L}_n < \bold{x} < \bold{R}_n$. It is important to point out that the count $a(\bold{x}) = N - d(\bold{x}) - u(\bold{x})$ is not the number of observations such that $\bold{R}_n < \bold{x}$. Finally, let $k(\bold{x})$ be the realized value of the actual count of observations such that $\bold{R}_n \leq \bold{x}$ and $E$ designate the event $d(\bold{x})\leq k(\bold{x})\leq d(\bold{x})+u(\bold{x})$.
Now we can estimate $F(\bold{x})$ by the value of the variable $p$ that maximizes the function
\[L(p; E) = \sum^{d(\bold{x})+u(\bold{x})}_{k=d(\bold{x})} {N \choose k}(p)^k (1-p)^{N-k}\]
subject to the constraint $0 \leq p \leq 1$. Consequently the estimator $\hat{F}$ of the unknown distribution function $F$ is given by
\[ \hat{F}(\bold{x}) = \left\{
\begin{array}{lll}
0 & \mbox{if} & d(\bold{x})=0 \mbox{ and } a(\bold{x})\geq 1 \\
\\
1 & \mbox{if} & a(\bold{x})=0 \mbox{ and } d(\bold{x})\geq 1 \\
\\
1/2 & \mbox{if} & u(\bold{x})=N \\
\\ \left(1+\sqrt[u\left(\bold{x}\right)+1]{\frac{a(\bold{x})(a(\bold{x})+1)\ldots(d(\bold{x})+u(\bold{x}))}{d(\bold{x})(d(\bold{x})+1)\ldots(d(\bold{x})+u(\bold{x}))}}\,\,\right)^{-1}&\mbox{o.w.}
\end{array}
\right. \]
We briefly consider once again a sample of univariate observations $X_1, \ldots, X_N$ with $X_n$ censored into an interval $(L_n, R_n]$ and assume that the random vectors $(X_n,\,L_n,\,R_n)$ are all i.i.d. according to some cdf $F_{XLR}$. The latter function provides a quantitative descsription of the censoring mechanism at play. By setting $\bold{X}_n = (X_n,\,L_n,\,R_n)$ and employing the estimation procedure just described we can construct an estimator $\hat{F}_{XLR}$ for the unknown function $F_{XLR}$ allowing us to estimate how the censoring mechanism operates.
\section{Kernel density estimation in 1 and 2 dimensions}
Consider a univariate random sample $Z_1,\,Z_2, \ldots, Z_N$ from some unknown pdf $f_Z$ and suppose that the corresponding observations are all exact. The kernel density estimate $\hat{f}_Z$ of $f_Z$ is defined as
\[ \hat{f}_Z(z) = \frac{1}{Nh}\,\sum_n K\left( \frac{z-z_n}{h} \right)\]
where $h$ is an appropriately chosen parameter. The rationale for such a construction is to place a ''bump'' of size $1/N$ centered over each one of the sample values $z_n$. The general shape of each bump is determined by the choice of the kernel function $K$ while its spread is controlled by the parameter $h$. All the bumps are set to be of equal size $1/N$ due to the i.i.d. nature of the observations. The size of the bump over $z_n$ can be also interpreted as the amount of probability assigned over the interval $(z_{n-1},\,z_n]$ by the empirical cdf $\hat{F}_z$ and is thus equal to $\hat{F}_Z(z_n)-\hat{F}_Z(z_{n-1})$.
We apply the reasoning from above to the case of a univariate random sample $X_1,\,X_2, \ldots, X_N$
from some unknown density function $f_X$ such that each $X_n$ is censored into some interval $(L_n,\,R_n]$. The set $G$ is reduced to the set $\left\{\tilde{X}_1,\,\tilde{X}_2,\ldots,\tilde{X}_I \right\}$ of unique element values of the set
\[ \left\{ L_1,\, L_2, \dots, L_N \right\} \cup \left\{ R_1,\, R_2, \dots, R_N \right\} \]
listed in increasing order. We proceed to define the function $w: G\rightarrow [0,1]$ by $w(\tilde{X}_1)= \hat{F}(\tilde{X}_1)$ and $w(\tilde{X}_i)=\hat{F}(\tilde{X}_i) - \hat{F}(\tilde{X}_{i-1})$ if $2\leq i \leq I$. Now we define the smoothed density estimator $\hat{f}_X$ as
\[ \hat{f}_X(x) = \frac{1}{h}\,\sum_i w(\tilde{X}_i)\, K\left( \frac{x-\tilde{X}_i}{h} \right)\]
Next we generalize the latter construction to the case of a random sample $\{(X_n, Y_n)\}$ of censored $2$-dimensional random vectors with unknown p.d.f $f_{XY}$. The set $G$ is given by $G = G^{(x)} \times G^{(y)}$ where
\begin{eqnarray*}
G^{(x)} &=& \{\, \tilde{X}_1, \tilde{X}_2, \ldots, \tilde{X}_I \,\}\\
G^{(y)} &=& \{\, \tilde{Y}_1, \tilde{Y}_2, \ldots, \tilde{Y}_J \,\}
\end{eqnarray*}
The definition of the function $w: G\rightarrow [0,1]$ is extended as follows: $w(\tilde{X}_i,\tilde{Y}_j)=0$ if $i=1$ or $j=1$. In all other cases $w(\tilde{X}_i,\tilde{Y}_j)$ equals the cummulative probability assigned by $\hat{F}$ over the interior of the rectangle in $R^2$ defined by the points $(\tilde{X}_{i-1},\tilde{Y}_{j-1})$, $(\tilde{X}_i,\tilde{Y}_{j-1})$, $(\tilde{X}_i,\tilde{Y}_j)$ and $(\tilde{X}_{i-1},\tilde{Y}_j)$ along with the line segments connecting $(\tilde{X}_i,\tilde{Y}_{j-1})$ with $(\tilde{X}_i,\tilde{Y}_j)$ and $(\tilde{X}_{i-1},\tilde{Y}_j)$ with $(\tilde{X}_i,\tilde{Y}_j)$. Consequently, the function value $w(\tilde{X}_i,\tilde{Y}_j)$, $2\leq i \leq I$, $2\leq j \leq J$, is given by
\[w(\tilde{X}_i,\tilde{Y}_j) = \hat{F}(\tilde{X}_i,\tilde{Y}_j) - \hat{F}(\tilde{X}_i,\tilde{Y}_{j-1}) - \hat{F}(\tilde{X}_{i-1},\tilde{Y}_j) + \hat{F}(\tilde{X}_{i-1},\tilde{Y}_{j-1}) \]
Pseudocode employing the recursive relationship from above to compute the weights $w(\tilde{X}_i,\tilde{Y}_j)$ is provided next:\\
\\
$\mbox{FOR } j=1:J$\\
$w(\tilde{X}_1,\tilde{Y}_j) = 0$\\
$\mbox{NEXT } j$\\
\\
$\mbox{FOR } i=1:I$\\
$w(\tilde{X}_i,\tilde{Y}_1) = 0$\\
$\mbox{NEXT } i$\\
\\
$\mbox{FOR } j=2:J$\\
$\mbox{FOR } i=2:I$\\
$w(\tilde{X}_i,\tilde{Y}_j) = \hat{F}(\tilde{X}_i,\tilde{Y}_j) - \hat{F}(\tilde{X}_i,\tilde{Y}_{j-1}) - \hat{F}(\tilde{X}_{i-1},\tilde{Y}_j) + \hat{F}(\tilde{X}_{i-1},\tilde{Y}_{j-1})$\\
$\mbox{NEXT } i$\\
$\mbox{NEXT } j$\\
Having developed a method for computing the weights $w(\tilde{x}_i,\tilde{y}_j)$ we are ready to present the expression for the smoothed density estimator $\hat{f}_{XY}(x, y)$:
\[ \hat{f}_{XY}(x, y) = \left(\frac{1}{h_x}\right) \left(\frac{1}{h_y}\right) \sum_{(i,j)} w(\tilde{X}_i,\,\tilde{Y}_j)\, K\left( \frac{x-\tilde{X}_i}{h_x} \right) K\left( \frac{y-\tilde{Y}_j}{h_y} \right)\]
\section{Kernel method in $M$ dimensions}
We will use $\tilde{\bold{X}}$ to designate an arbitrary element $\left(\tilde{X}_{i_1},\, \tilde{X}_{i_2}, \ldots, \, \tilde{X}_{i_M} \right)$ of the grid $G$. Furthermore, given any vector $\tilde{\bold{X}} \in G$ such that $i_m\geq 2$ for $\forall m$ we will define the vector
\[ \tilde{\bold{X}}' = \left(\tilde{X}_{i_1-1},\, \tilde{X}_{i_2-1}, \ldots, \, \tilde{X}_{i_M-1} \right)\in G \]
Let $\Omega(\bold{x})$ be the set of all hyperplanes passing through $\bold{x}$ and parallel to the coordinate planes. For example, $\Omega(\tilde{\bold{X}})$ is the set of all hyperplanes passing through $\tilde{\bold{X}}$ and parallel to the coordinate planes. Define the function $w: G\rightarrow [0,1]$ as follows: $w( \tilde{\bold{X}} ) = \hat{F}( \tilde{\bold{X}} ) = 0$ if there exists a component $\tilde{X}_{i_m}$ of $\tilde{\bold{X}}$ such that $i_m=1$. In the case when $i_m\geq 2$ for $\forall m$ the value of $w( \tilde{\bold{X}} )$ is given by the cummulative probability assigned by $\hat{F}$ over the hyperrectangle in $R^M$ bounded by the hyperplanes in $\Omega(\tilde{\bold{X}}')$ and $\Omega(\tilde{\bold{X}})$ but excluding the points lying on the hyperplanes in $\Omega(\tilde{\bold{X}}')$.
Let $\bold{h} = (h_1,\,h_2,\ldots,h_M)$. The smoothed function estimator is given by
\[ \hat{f}(\bold{x}) = \left(\prod_m \frac{1}{h_m}\right) \sum_{\tilde{\bold{x}}} w(\tilde{\bold{x}}) K\left(\bold{x};\tilde{\bold{x}},\bold{h}\right) \]
where
\[ K\left(\bold{x};\tilde{\bold{x}},\bold{h}\right) = \prod_m K\left(\frac{x_m - \tilde{x}_{i_m}}{h_m} \right) \]
\section{A loss function for computing the optimal bandwidth}
Consider once again a univariate random sample $Z_1,\,Z_2, \ldots, Z_N$ from some unknown pdf $f_Z$ and the kernel density estimator
\[ \hat{f}_Z(z) = \hat{f}_Z(z; h) = \frac{1}{N h}\,\sum_n K\left( \frac{z-Z_n}{h} \right)\]
The bandwidth $h$ will be treated as a variable for the remainder of the section. Also, to simplify notation whenever no ambiguity arises we will distinguish density functions by their argument only and drop the subscripting random variable. For example, $f(z)$ will represent $f_Z(z)$. In addition, we will use a subscript ``$-n$'' to indicate that a quantity has been derived based on the subset of the original random sample obtained after removing the $n$-th observation. For example, $\hat{f}_{-n}(z)$ is the kernel density estimator for $f(z)$ calculated after removing $Z_n$ from the original sample.
The integrated square error is defined as
\[ \int \left[\hat{f}(z)-f(z) \right]^2dz = \int \hat{f}(z)^2dz - 2\int\hat{f}(z)f(z)dz + \int f(z)^2dz \]
and the value of $h$ minimizing the risk function $R(h)$ given by
\begin{eqnarray*}
R(h) &=& E\left\{ \int \left[\hat{f}(z)-f(z) \right]^2dz \right\} \\
R(h) &=& E\left\{\int \hat{f}(z)^2dz \right\} - 2\,E\left\{\int\hat{f}(z)f(z)dz \right\} + \int f(z)^2dz
\end{eqnarray*}
is generally viewed as the optimal choice for the value of $h$ in $\hat{f}_Z(z; h)$. The term $\int f(z)^2dz$ is independent of $h$ and as a result we need to minimize
\[ E\left\{ \int \hat{f}(z)^2dz \right\} - 2\,E\left\{ \int\hat{f}(z)f(z)dz \right\} \]
The latter goal, however, is unachievable since the density $f_Z$ is unknown.
In reality we seek to minimize the score function
\[ M_0(h) = \int \hat{f}(z)^2dz - \frac{2}{N}\sum_n f_{-n}(Z_n) \]
for two reasons. It is straightforward to demonstrate that
\[ E\left\{ \frac{1}{n}\sum_n f_{-n}(Z_n) \right\} = E\left\{ \int\hat{f}(z)f(z)dz \right\} \]
which immediately implies that
\[E\left\{ M_0(h) \right\} = E\left\{ \int \hat{f}(z)^2dz \right\} - 2\,E\left\{ \int\hat{f}(z)f(z)dz \right\} \]
In addition, as stated by Silverman \cite{r2} ``Assuming that the minimizer of $M_0(h)$ is close to the minimizer of $E\{M_0(h)\}$ indicates why we might hope that minimizing $M_0$ gives a good choice of smoothing parameter.''
Now we move on to motivate and introduce a score function $\tilde{M}_0(h)$ that mimics the form of $M_0(h)$ and can be used in the presence of censoring. We begin by defining the random variables
\begin{eqnarray*}
\tilde{L}_n &=& \mbox{max}\{-\infty,\,L_n\}\\
\tilde{R}_n &=& \mbox{min}\{R_n,\,+\infty\}\\
V_n &=& \frac{1}{2}(\tilde{L}_n+\tilde{R}_n)
\end{eqnarray*}
If we make the assumption that the probability distribution functions $g$ of $V_n$ and $f$ of $X_n$ are approximately equal, i.e $g(v) \approx f(v)$, then we have that
\begin{eqnarray*}
E\left\{\frac{1}{N} \sum \hat{f}_{-n}(V_n) \right\} &=& \frac{1}{N}(N) E\left\{ \hat{f}_{-1}(V_1) \right\}\\
\\
&=& E\left\{ \hat{f}_{-1}(V_1) \right\}\\
\\
&=& E\left\{ \int\hat{f}_{-1}(v)g(v)dv \right\}\\
\\
&\approx& E\left\{ \int\hat{f}_{-1}(v)f(v)dv \right\}\\
\\
&=& E\left\{ \int\hat{f}_{-1}(x)f(x)dx \right\}\\
\end{eqnarray*}
Since the expected values $E\left\{ \int\hat{f}_{-1}(x)f_X(x)dx \right\}$ and $E\left\{ \int\hat{f}(x)f_X(x)dx \right\}$ converge asymptotically we can conclude that for large samples
\[ E\left\{\frac{1}{N} \sum \hat{f}_{-n}(V_n) \right\} \approx E\left\{ \int\hat{f}_{-1}(x)f(x)dx \right\} \approx E\left\{ \int\hat{f}(x)f(x)dx \right\} \]
Consequently, we define $\tilde{M}_0(h)$ as
\[ \tilde{M}_0(h) = \int \hat{f}(x)^2dx - \frac{2}{N} \sum \hat{f}_{-n}(V_n)\]
In $M\geq 2$ dimensions we define the random variables
\begin{eqnarray*}
\tilde{L}_{nm} &=& \mbox{max}\{-\infty,\,L_{nm}\}\\
\tilde{R}_{nm} &=& \mbox{min}\{R_{nm},\,+\infty\}\\
V_{nm} &=& \frac{1}{2}(\tilde{L}_{nm}+\tilde{R}_{nm})
\end{eqnarray*}
and the random vector $\bold{V}_n = (V_{n1},\, V_{n2}, \ldots, V_{nM})$. Under the assumption that the probability distribution functions $g$ of $\bold{V}_n$ and $f$ of $\bold{X}_n$ are approximately equal, i.e $g(\bold{v}) \approx f(\bold{v})$, and based on identical reasoning we generalize the definition of $\tilde{M}_0(\bold{h})$ as follows:
\[ \tilde{M}_0(\bold{h}) = \int \hat{f}(\bold{x})^2d\bold{x} - \frac{2}{N} \sum \hat{f}_{-n}(\bold{V}_n)\]
\section{Nadaraya Watson regression with censored data}
In regression analysis the goal is to estimate the expected value $E\left\{Y|\bold{X}=\bold{x}\right\}$ based on a random sample $\left\{(\bold{X}_n,\,Y_n) \right\}$ from some unknown p.d.f. $f$ where $\bold{X}_n$ is an $M$-dimensional vector of explanatory variables. Nadaraya and Watson \cite{r3, r4} have proposed a non-parametric estimator for $E\left\{Y|\bold{X}=\bold{x}\right\}$ derived from the kernel density estimator for $f$ in the case when all sample observations are exact. We employ the newly developed censoring kernel density estimator
\[ \hat{f}(\bold{x}) = \left(\prod_m \frac{1}{h_m}\right) \sum_{\tilde{\bold{x}}} w(\tilde{\bold{x}}) K\left(\bold{x};\tilde{\bold{x}},\bold{h}\right) \]
and an identical pattern of reasoning to adapt the Nadaraya-Watson estimator for use with censored data.
In $1+1$ dimensions the censoring kernel density estimator can be written as
\[ \hat{f}(x,y) = \sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j)\frac{1}{h_x\, h_y} K\left(\frac{x-\tilde{x}_i}{h_x}\right)K\left(\frac{y-\tilde{y}_j}{h_y}\right) \]
Consequently
\begin{eqnarray*}
\hat{f}(x) &=& \int \hat{f}(x,y) dy\\
\\
&=& \sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j)\frac{1}{h_x\, h_y} K\left(\frac{x-\tilde{x}_i}{h_x}\right) \int K\left(\frac{y-\tilde{y}_j}{h_y}\right)dy\\
\\
&=& \sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j)\frac{1}{h_x\, h_y} K\left(\frac{x-\tilde{x}_i}{h_x}\right)h_y\\
\\
&=& \frac{1}{h_x}\, \sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j) K\left(\frac{x-\tilde{x}_i}{h_x}\right)
\end{eqnarray*}
and
\begin{eqnarray*}
\int y\hat{f}(x,y) dy &=& \sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j)\frac{1}{h_x} K\left(\frac{x-\tilde{x}_i}{h_x}\right) \int y\, \frac{1}{h_y} K\left(\frac{y-\tilde{y}_j}{h_y}\right)dy\\
\\
&=& \frac{1}{h_x}\,\sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j) K\left(\frac{x-\tilde{x}_i}{h_x}\right)\tilde{y}_j
\end{eqnarray*}
Now we define the estimator $E\left\{Y|X=x\right\}$ as follows:
\[ E\left\{Y|X=x\right\} = \frac{\int y\hat{f}(x,y) dy}{\hat{f}(x)} = \frac{\sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j) K\left(\frac{x-\tilde{x}_i}{h_x}\right)\tilde{y}_j}{\sum_{(i,j)} w(\tilde{x}_i, \tilde{y}_j) K\left(\frac{x-\tilde{x}_i}{h_x}\right)} \]
In $(M+1)$ dimensions the same reasoning leads us to define the estimator $E\left\{Y|\bold{X}=\bold{x}\right\}$ as
\[ E\left\{Y|\bold{X}=\bold{x}\right\} = \frac{\int y\hat{f}(\bold{x},y) dy}{\hat{f}(\bold{x})} = \frac{\sum_{(\tilde{\bold{x}},y_j)} w(\tilde{\bold{x}}, \tilde{y}_j) K\left(\bold{x};\tilde{\bold{x}},\bold{h}\right)\tilde{y}_j} {\sum_{(\tilde{\bold{x}},y_j)} w(\tilde{\bold{x}}, \tilde{y}_j) K\left(\bold{x};\tilde{\bold{x}},\bold{h}\right)} \]
\section{Parameter estimation for a multinomial distribution in the presence of censoring}
Let $c_1$ and $c_2$ be the respective observed numbers of outcomes of type 1 and type 2 in a binomial experiment with $N$ trials, $u = N-c_1-c_2\geq 1$ number of trials with unknown outcomes and probability $\pi$ of a single trial being of type 1. Let $N_1$ and $N_2$ designate the actual counts of type 1 and type 2. Consequently $N_1$ and $N_2$ are censored such that $(N_1,\,N_2)\in S_2$ where the set $S_2$ is defined by
\[ S_2 = \{(l_1, l_2) \,|\, l_1,l_2 \mbox{ are non-negative integers},\, l_1\geq c_1,\, l_2\geq c_2,\, l_1+l_2=N\} \]
If $E$ designates the event $(N_1,\,N_2)\in S$ then the likelihood of observing $E$ is given by
\[L(\pi; E) = \sum_{(n_1,n_2)\in S_2} \frac{N!}{n_1!\,n_2!}\,(\pi)^{n_1}\,(1-\pi)^{n_2} = \sum^{c_1+u}_{n_1=c_1} {N \choose n_1}(\pi)^{n_1} (1-\pi)^{N-n_1}\]
As already derived, the value $\hat{\pi}$ of $p$ that maximizes the function
\[L(p; E) = \sum^{c_1+u}_{n_1=c_1} {N \choose n_1}(p)^{n_1} (1-p)^{N-n_1}\]
subject to the constraint $0 \leq p \leq 1$ is given by
\[ \hat{\pi} = \left\{
\begin{array}{lll}
0 & \mbox{if} & c_1=0 \mbox{ and } c_2\geq 1 \\
\\
1 & \mbox{if} & c_2=0 \mbox{ and } c_1\geq 1 \\
\\
1/2 & \mbox{if} & u=N \\
\\ \left(1+\sqrt[u+1]{\frac{c_2 (c_2+1)\ldots(c_2+u)}{c_1 (c_1+1)\ldots(c_1+u)}}\,\right)^{-1}&\mbox{o.w.}
\end{array}
\right. \]
The treatment of an multinomial experiment with $N$ trials, $M$ possible outcome types and probabilities $\pi_1, \pi_2, \ldots, \pi_M$ of each outcome type is based on the same reasoning. We use $c_1, c_2, \ldots, c_M$ to designate the observed counts of each type and $N_1, N_2, \ldots, N_M$ to designate the actual and possibly censored outcome counts. Suppose $u = N - \sum_m c_m \geq 1$ and define the vectors
\begin{eqnarray*}
\bold{p} &=& (p_1,\,p_2,\ldots,p_M)\\
\bold{c} &=& (c_1,\,c_2,\ldots,c_M)\\
\bold{n} &=& (n_1,\,n_2,\ldots,n_M)\\
\bold{N} &=& (N_1,\,N_2,\ldots,N_M)
\end{eqnarray*}
The definition of the set $S_2$ generalizes to
\[ S_M = \{(l_1, l_2, \ldots, l_M) \,|\, l_{m}\mbox{ is a non-negative integer},\, l_{m}\geq c_{m},\, \sum_m l_m=N \} \]
and accordingly $E$ is redefined to be the event $(N_1,\,N_2, \ldots, N_M)\in S_M$. The likelihood of $E$ as a function of $\bold{p}$ is given by
\[L(\bold{p}; E) = \sum_{\bold{n}\in S_M} \frac{N!}{n_1!\,n_2!\ldots n_M!}\,(p_1)^{n_1}\,(p_2)^{n_2}\ldots \,(p_M)^{n_M}\]
An approximate solution to the resulting estimation problem can be constructed as follows. If $\hat{p}_m$ is the value of the variable $p_m$ that maximizes the function
\[ \sum^{c_m+u}_{n_m=c_m} {N \choose n_m}(p_m)^{n_m} (1-p_m)^{N-n_m} \]
then we could employ
\[\hat{\pi}^*_m = \frac{\hat{p}_m}{\hat{p}_1+\hat{p}_2+\ldots+\hat{p}_M}\]
as an estimator for the unknown probability $\pi_m$.
Next we consider a trinomial $(M=3)$ experiment such that $u_{12}$ trials are of type $1$ or type $2$ and $u_{23}$ are of type $2$ or type $3$ and define
\begin{eqnarray*}
u_1 &=& u_{12}\\
u_2 &=& \mbox{min}\{ N - c_1 -c_2 - c_3,\, u_{12}+u_{23} \}\\
u_3 &=& u_{23}
\end{eqnarray*}
Let $\hat{p}_m$ be the value of the variable $p_m$ that maximizes the function
\[ \sum^{c_m+u_m}_{n_m=c_m} {N \choose n_m}(p_m)^{n_m} (1-p_m)^{N-n_m} \]
and
\[\hat{\pi}^*_m = \frac{\hat{p}_m}{\hat{p}_1+\hat{p}_2+\ldots+\hat{p}_M}\]
The quantities $\hat{\pi}^*_1$, $\hat{\pi}^*_2$ and $\hat{\pi}^*_3$ can be used to estimate the unknown probabilities $\pi_1$, $\pi_2$ and $\pi_3$. Generalizing to the case of $M$ possible outcomes in the presence of partial censoring is straightforward. Let $u_m$ be the maximum possible number of censored outcomes of type $m$ and assume that $1, 2, \ldots, u_m$ are all possible counts for the number of unobserved outcomes of type $m$. Consequently $\hat{\pi}^*_m$ is a potential estimator for $\pi_m$.
So far we have been constructing likelihood functions without making assumptions or having the benefit of prior knowledge about the nature of the censoring mechanism. Let $q_m$ be the conditional probability of observing an outcome ot type $m$ and $\bold{q}=(q_1,\,q_2,\ldots,q_M)$. For example, let us consider a binomial ($M=2$) experiment with known parameters $q_1$ and $q_2$. The probability of not being able to observe the outcome of a single trial $X_n$ is given by $(1-q_1)p_1+(1-q_2)p_2 = (p_1+p_2)-p_1q_1-p_2q_2$. Consequently the likelihood of observing $c_1$ outcomes of type $1$, $c_2$ outcomes of type $2$ and $u = N-c_1-c_2$ outcomes of unknown type is
\[L(\bold{p}, \bold{q};\bold{c},\,u) = \frac{N!}{c_1!\,c_2!\,u!}\,(p_1\,q_1)^{c_1}\,(p_2q_2)^{c_2}\left[(1-q_1)\,p_1+(1-q_2)\,p_2\right]^u\]
Generalizing is trivial:
\[L(\bold{p}, \bold{q};\bold{c},\,u) = \frac{N!}{c_1!\,c_2! \ldots c_M!\,u!}\,\prod_m{(p_m\,q_m)^{c_1}}\,\left[\sum_m{(1-q_m)\,p_m}\right]^u\]
where $u = N - \sum_m c_m$.
Finally we turn our attention to a binomial experiment such that $q_1$ remains unknown but $q_2$ is known. The outcome $x_n$ of a single trial $X_n$ can be classified in exactly one of the following four categories: observed of type 1, observed of type 2, unobserved of type 1 and unobserved of type 2. Let $\tilde{N}_1$ designate the number of censored outcomes of type 1, $\tilde{N}_2$ designate the number of censored outcomes of type 2 and the set $\tilde{S}_2$ be defined as
\[ \tilde{S}_2 = \{(l_1, l_2) \,|\, l_1,\,l_2\mbox{ are non-negative integers},\, l_1+l_2=N-c_1-c_2 \} \]
The likelihood of the event $\tilde{E}=\,$``$(\tilde{N}_1,\,N_2)\in \tilde{S}_2$'' is given by
\[ L(\bold{\pi}, \bold{q}; \tilde{E}) = \sum_{(\tilde{n}_1,\tilde{n}_2)} \frac{N!}{c_1!\,c_2!\,\tilde{n}_1!\,\tilde{n}_2!}\,(\pi_1q_1)^{c_1}\,(\pi_2q_2)^{c_2}\left[(1-q_1)\pi_1\right]^{\tilde{n}_1}\left[(1-q_2)\pi_2\right]^{\tilde{n}_2} \]
where the summation index $(\tilde{n}_1,\tilde{n}_2)$ spans the set $\tilde{S}_2$. Consequently we seek to maximize the function
\[ L(\bold{p},\,q'_2; E) = \sum_{(\tilde{n}_1,\tilde{n}_2)} \frac{N!}{c_1!\,c_2!\,\tilde{n}_1!\,\tilde{n}_2!}\,(p_1q_1)^{c_1}\,(p_2q_2')^{c_2}\left[(1-q_1)p_1\right]^{\tilde{n}_1}\left[(1-q_2')p_2\right]^{\tilde{n}_2} \]
subject to the constraints $p_1+p_2=1$ and $0\leq q_2' \leq 1$.
\section{Analysis of contingency tables with incomplete counts}
Since each cell in an $I\times J$ contingency table can be uniquely associated with an ordered pair $(i,j)$ the set of ordered pairs $\{(i,j)\}$ constitutes the space of possible outcomes for a sample random variable $X_n$. Define the probabilities $\pi_{ij}$, $q_{ij}$ and $\alpha_{ij}$ as
\begin{eqnarray*}
\pi_{ij} &=& \mbox{Prob}\left\{ X_n = (i,j) \right\}\\
q_{ij} &=& \mbox{Prob}\left\{ X_n \mbox{ is observed}\,|\,X_n=(i,j) \right\}\\
\alpha_{ij} &=& \mbox{Prob}\left\{ X_n=(i,j) \mbox{ and }X_n \mbox{ is observed} \right\} = \pi_{ij} q_{ij}
\end{eqnarray*}
Furthermore, let $c_{ij}$ and $N_{ij}$ be the respective observed and actual counts in cell $(i,j)$. We can quantify the effect of the censoring mechanism by observing that the ratio $\hat{\alpha}_{ij} = \frac{c_{ij}}{N}$ constitutes an MLE for the joint probability $\alpha_{ij}$ and using the plug-in principle within the equation $\alpha_{ij} = \pi_{ij} q_{ij}$ to obtain the estimator $\hat{q}_{ij} = \frac{c_{ij}}{\hat{\pi}_{ij}N}$ for the unknown probability $q_{ij}$.
The actual count $N_{ij}$ may be unknown due to the censoring mechanism. From the definitions follows that $c_{ij} = N_{ij}$ if outcomes of type $(i,j)$ are not subject to censoring and $c_{ij} \leq N_{ij}$ otherwise. Finally, let us use $N_j = \sum_i N_{ij}$ to designate the $j$-th column total and in the case when $N_j$ is known let $u_j = N_j - \sum_i c_{ij}$ designate the number of sample outcomes censored into the $j$-th column.
Consider the special case of a $2\times 2$ ($I=2, J=2$) contingency table and the null hypothesis
\[ H_0:\,\mbox{Prob}\left\{ X_n = (1,1)\,|\, X_n \in \{(1,1),(2,1)\} \right\} = \mbox{Prob}\left\{ X_n = (1,2)\,|\, X_n \in \{(1,2),(2,2)\} \right\} \]
which can be rewritten as
\[ H_0:\, \frac{\pi_{11}}{\pi_{11}+\pi_{21}} = \frac{\pi_{12}}{\pi_{12}+\pi_{22}} \]
Assuming $H_0$ in an estimation problem amounts to introducing the constraint
\[ \frac{p_{11}}{p_{11}+p_{21}} = \frac{p_{12}}{p_{12}+p_{22}} \]
where $p_{ij}$ is the variable associated with the unknown cell probability $\pi_{ij}$. In the special case of predetermined column totals $N_1$ and $N_2$ we have that $\pi_{11}+\pi_{12}=1$ as well as $\pi_{12}+\pi_{22}=1$. Consequently, the null hypothesis is reduced to $H_0:\, \pi_{11}=\pi_{12}=\pi$ and accordingly the null constraint becomes $p_{11}=p_{22}=p$.
Before turning our attention to three examples of censored $2\times 2$ contingency tables we introduce some additional notation:
\begin{eqnarray*}
&S& = \{(l_{11},\,l_{21},\,l_{12},\,l_{22}) \,|\, l_{ij}\mbox{ is a non-negative integer},\, l_{ij}\geq c_{ij},\, \sum_{(i,j)} l_{ij}=N \}\\
&\bold{N}& = (N_{11},\,N_{21},\,N_{12},\,N_{22})\\
&\bold{n}& = (n_{11},\,n_{21},\,n_{12},\,n_{22})\\
&\bold{c}& = (c_{11},\,c_{21},\,c_{12},\,c_{22})\\
&\bold{p}& = (p_{11},\,p_{21},\,p_{12},\,p_{22})\\
&\bar{\bold{c}}& = (\bar{c}_{11},\,\bar{c}_{21},\,\bar{c}_{12},\,\bar{c}_{22})
\end{eqnarray*}
In each example we construct the likelihood necessary to derive a set of estimators $\{\hat{\pi}_{ij}\}$ for the elements of $\{\pi_{ij}\}$. A superscript ``$(0)$'' will be used to label quantities derived under $H_0$. For example, $\hat{\pi}^{(0)}_{ij}$ is the null esimator for $\pi_{ij}$.
\subsection{Example 1}
Suppose that $N_1$ and $N_2$ are predetermined by the experimenter, the counts $N_{11}$ and $N_{21}$ are exact implying $u_1=0$ while the counts $N_{12}$ are $N_{22}$ are censored implying $u_2\geq 1$. Let $E_1$ designate the event ``$\bold{N}\in S$ and $N_{11}+N_{21}=N_1$ and $N_{12}+N_{22}=N_2$''. The likelihood of observing $E_1$ is given by
\[ L(\bold{p}; E_1) = L_1(p_{11})\,L_2(p_{12})\]
where
\begin{eqnarray*}
L_1(p_{11}) &=& {N_1 \choose N_{11}}(p_{11})^{N_{11}} (1-p_{11})^{N_1-N_{11}} \\
\\
L_1(p_{12}) &=& \sum^{c_{12}+u_2}_{n_{12}=c_{12}} {N_2 \choose n_{12}}(p_{12})^{n_{12}} (1-p_{12})^{N_2-n_{12}}
\end{eqnarray*}
Since the column totals $N_1$ and $N_2$ are fixed and known in advance, under $H_0$ the likelihood function needs to be modified by setting $p = p_{11} = p_{12}$:
\begin{eqnarray*}
L(\bold{p}; E, H_0) &=& \sum^{c_{12}+u_2}_{n_{12}=c_{12}} {N_1 \choose N_{11}}(p)^{N_{11}} (1-p)^{N_1-N_{11}} \, {N_2 \choose n_{12}}(p)^{n_{12}} (1-p)^{N_2-n_{12}} \\
\\
L(\bold{p}; E, H_0) &=& \sum^{c_{12}+u_2}_{n_{12}=c_{12}} {N_1 \choose N_{11}}{N_2 \choose n_{12}}(p)^{N_{11}+n_{12}} (1-p)^{N-N_{11}-n_{12}}
\end{eqnarray*}
Let $\bold{t} = (\bar{N}_{11},\,\bar{c}_{12},\,\bar{c}_{22},\,\bar{u}_2)$ be a particular vector of counts for the contingency table. Then the probability of observing $\bold{t}$ is given by
\[ \mbox{Prob}\{\bold{t}\} = P_1(\bar{N}_{11})\,P_2(\bar{c}_{12},\,\bar{c}_{22},\,\bar{u}_2) \]
where
\begin{eqnarray*}
P_1(\bar{N}_{11}) &=& \frac{N_1}{\bar{N}_{11}!\,\bar{N}_{21}!}\,(\pi_{11})^{\bar{N}_{11}}\,(\pi_{21})^{\bar{N}_{21}}\\
\\
P_2(\bar{c}_{12},\,\bar{c}_{22},\,\bar{u}_2) &=& \frac{N_2}{\bar{c}_{12}!\,\bar{c}_{22}!\,\bar{u}_2!}\,(\alpha_{12})^{\bar{c}_{12}}\,(\alpha_{22})^{\bar{c}_{22}}\,(1-\alpha_{12}-\alpha_{22})^{\bar{u}_2}
\end{eqnarray*}
We can estimate $\mbox{Prob}\{\bold{t}\}$ by using $\hat{\pi}_{11},\, \hat{\pi}_{21},\, \hat{\alpha}_{12}$ and $\hat{\alpha}_{22}$ for the unknown probabilities $\pi_{11},\, \pi_{21},\, \alpha_{12}$ and $\alpha_{22}$. Under $H_0$ we estimate $\mbox{Prob}\{\bold{t}\}$ by employing the appropriate null estimators $\hat{\pi}^{(0)}_{11}$ and $\hat{\pi}^{(0)}_{21}$ as opposed to $\hat{\pi}_{11}$ and $\hat{\pi}_{21}$.
\subsection{Example 2}
Suppose $N_1$ and $N_2$ are predetermined by the experimenter and the counts $N_{11}, N_{21}, N_{12}, N_{22}$ are all unobserved. We use $E_2$ designate the event ``$\bold{N}\in S$ and $N_{11}+N_{21}=N_1$ and $N_{12}+N_{22}=N_2$''. The likelihood of $E_2$ is given by
\[ L(\bold{p}; E_2) = L_1(p_{11})\,L_2(p_{12})\]
where
\begin{eqnarray*}
L_1(p_{11}) &=& \sum^{c_{11}+u_1}_{n_{11}=c_{11}} {N_1 \choose n_{11}}(p_{11})^{n_{11}} (1-p_{11})^{N_1-n_{11}} \\
\\
L_1(p_{12}) &=& \sum^{c_{12}+u_2}_{n_{12}=c_{12}} {N_2 \choose n_{12}}(p_{12})^{n_{12}} (1-p_{12})^{N_2-n_{12}}
\end{eqnarray*}
The two factors $L_1(p_{11}; E)$ and $L_2(p_{12}; E)$ can be maximized independently if no further assumptions are made. Under the null constraint $p = p_{11} = p_{12}$ the likelihood $L(\bold{p}; E_2)$ is modified as follows:
\begin{eqnarray*}
L(\bold{p}; E_2, H_0) &=& \sum_{(n_{11},\,n_{12})} {N_1 \choose n_{11}}(p)^{n_{11}} (1-p)^{N_1-n_{11}} \, {N_2 \choose n_{12}}(p)^{n_{12}} (1-p)^{N_2-n_{12}} \\
\\
L(\bold{p}; E_2, H_0) &=& \sum_{(n_{11},\,n_{12})} {N_1 \choose n_{11}}{N_2 \choose n_{12}}(p)^{n_{11}+n_{12}}\, (1-p)^{N-n_{11}-n_{12}}
\end{eqnarray*}
where $(n_{11},\,n_{12}) \in S$ and $n_{11}+n_{21}=N_1$ and $n_{12}+n_{22}=N_2$.
The probability of a particular contingency table configuration is given by
\[ \mbox{Prob}\{\bar{\bold{c}},\,\bar{u}_1,\,\bar{u}_2 \} = P_1(\bar{c}_{11},\,\bar{c}_{21},\,\bar{u}_1)\,P_2(\bar{c}_{12},\,\bar{c}_{22},\,\bar{u}_2) \]
with
\begin{eqnarray*}
P_1(\bar{c}_{11},\,\bar{c}_{21},\,\bar{u}_1) &=& \frac{N_1}{\bar{c}_{11}!\,\bar{c}_{21}!\,\bar{u}_1!}\,(\alpha_{11})^{\bar{c}_{11}}\,(\alpha_{21})^{\bar{c}_{21}}\,(1-\alpha_{11}-\alpha_{21})^{\bar{u}_1}\\
\\
P_2(\bar{c}_{12},\,\bar{c}_{22},\,\bar{u}_2) &=& \frac{N_2}{\bar{c}_{12}!\,\bar{c}_{22}!\,\bar{u}_2!}\,(\alpha_{12})^{\bar{c}_{12}}\,(\alpha_{22})^{\bar{c}_{22}}\,(1-\alpha_{12}-\alpha_{22})^{\bar{u}_2}
\end{eqnarray*}
The estimators for $\alpha_{ij}$ remain unchanged under $H_0$ unless additional assumptions are made regarding the nature of the censoring mechanism.
\subsection{Example 3}
Suppose that $N_{11}, N_{21}, N_{12}, N_{22}$ as well as the column totals $N_1$ and $N_2$ are all unknown. Let
$u = N - (c_{11}+c_{21}+c_{12}+c_{22})$ and $E_3$ designate the event ``$\bold{N}\in S$''. The estimators $\hat{\pi}_{ij}$ maximize the likelihood
\[L(\bold{p}; E_3) = \sum_{\bold{n}\in S} \frac{N!}{n_{11}!\,n_{21}!\,n_{12}!\,n_{22}!}\,(p_{11})^{n_{11}}\,(p_{21})^{n_{21}} \,(p_{12})^{n_{12}}\,(p_{22})^{n_{22}}\]
By enforcing the null constraint the above likelihood is reduced to
\begin{eqnarray*}
L(\bold{p}; E_3, H_0) &=& \sum_{\bold{n}\in S} \frac{N!}{n_{11}!\,n_{21}!\,n_{12}!\,n_{22}!}\,(p)^{n_{11}}\,(1-p)^{n_{21}} \,(p)^{n_{12}}\,(1-p)^{n_{22}}
\\
L(\bold{p}; E_3, H_0) &=& \sum_{\bold{n}\in S} \frac{N!}{n_{11}!\,n_{21}!\,n_{12}!\,n_{22}!}\,(p)^{n_{11}+n_{12}}\,(1-p)^{n_{21}+n_{22}}
\end{eqnarray*}
The probability of a particular contingency table configuration is given by
\[ \mbox{Prob}\{ \bar{\bold{c}},\,\bar{u} \} = \frac{N}{\bar{c}_{11}!\,\bar{c}_{21}!\,\bar{c}_{12}!\,\bar{c}_{22}!\,\bar{u}!}\, P_1(\bar{\bold{c}})\,P_2(u) \]
where
\begin{eqnarray*}
P_1(\bar{\bold{c}}) &=& (\alpha_{11})^{\bar{c}_{11}}\,(\alpha_{21})^{\bar{c}_{21}}\,(\alpha_{12})^{\bar{c}_{12}}\,(\alpha_{22})^{\bar{c}_{22}}\\
\\
P_2(u) &=& (1-\alpha_{11}-\alpha_{21}-\alpha_{12}-\alpha_{22})^{\bar{u}}
\end{eqnarray*}
Assuming $H_0$ does not modify the estimate for $\mbox{Prob}\{ \bar{\bold{c}},\,\bar{u} \}$.\\
Extending the ideas presented in this section to the construction of appropriate likelihood functions for contingency tables with $I\geq 2$ rows and $J\geq 2$ columns in the presence of a censoring mechanism should be trivial in most cases. Solving the resulting optimizataion problems, however, may be far from straightforward.\\
|
\section{Introduction}
In the Integral Rectangle Packing (IRP) problem, we are given a set $J=\{1,2,\cdots,n\}$ of $n$ integral rectangles, each having width $ w_j$ and height $h_j$, and an integral rectangular container of width $W$ and height $H$. The problem is to determine whether all rectangles can be orthogonally placed into the container without overlapping. If the answer is yes, then we should present a non-overlapping packing pattern. In this paper, we assume that: (1) $w_j$, $h_j$, $W$ and $H$ are positive integers, and each vertex of each rectangle must be at integral point in the container. (2) Rectangles are rotatable, i.e., each rectangle can be horizontally or vertically placed into the container.
The rectangle packing problem arises in many industrial applications, such as cutting wood, glass, paper and steel in manufacturing, packing goods in transportation, arranging articles and advertisements in publishing. Various algorithms have been proposed to solve this problem. They can be divided into three categories: approximate algorithms, heuristic algorithms and exact algorithms \cite{hopper2001,lodi2002}.
Most algorithms solve the RP problem by successively placing a new rectangle into the container. Then a basic problem arises: where to place a new rectangle when the container is already partially occupied by some previously placed rectangles? To handle this problem, people have proposed several kinds of placement heuristics which specifies admissible positions for a rectangle \cite{baker1980, berkey1987, burke2004, huang2011, huang2007}.
However, for a specific placement heuristic, there exists an important question: when there exist feasible solutions, can we achieve one by successively placing a rectangle into the container using this placement heuristic? If the answer is yes, then we say the placement heuristic is \textbf{complete}; otherwise, \textbf{incomplete}. If a placement heuristic is incomplete, then any algorithm based on it is foredoomed to fail on some instances. For example, \citet{baker1980} have proved that the Bottom-Left heuristic, which places a rectangle onto the lowest possible position in the container and left-justify it, is incomplete. They found an instance for which any feasible solution can not be achieved using the Bottom-Left heuristic, no matter what ordering of the rectangles is used. \citet{martello2000} proposed a placement heuristic and developed an exact algorithm for the three dimensional bin packing problem. Later, \citet{boef2005} found that some instances can not be solved using the placement heuristic proposed by \citet{martello2000}.
It is usually very difficult to prove a placement heuristic's completeness or incompleteness. Nevertheless, for a special case of the RP problem, the 2D rectangular perfect packing problem, \citet{lesh2004} have shown that the Bottom-Left heuristic is complete. They presented the following theorem: \textit{For every perfect packing, there is a permutation of the rectangles that yields that packing using the Bottom-Left heuristic.} An efficient branch and bound algorithm is also developed based on this theorem. Besides this result, we have not found other paper in literature proving a certain placement heuristic's completeness.
For the general IRP problem, this paper formulates a placement heuristic and proves its completeness. The following corner occupying theorem is presented that \textit{ if it is possible to orthogonally place $n$ arbitrarily given integral rectangles into an integral rectangular container without overlapping, then we can achieve a feasible packing by successively placing an integral rectangle onto a bottom-left corner in the container.} This theorem lays a solid foundation for understanding many efficient and exact algorithms for solving the RP problem\cite{christofides1977,hadji1995, kenmochi2009}. It might be possible to develop new efficient and effective heuristic algorithms based on this theorem.
The rest of the paper is organized as follows. Section 2 presents several notations and definitions. Section 3 proves the corner occupying theorem. Finally, Section 4 presents a counterexample to show why the proof of corner occupying theorem can not be directly extended to the three dimensional case.
\section{Notations and definitions in integral rectangle packing problem}
We designate the bottom-left corner point of the container as the origin of the $xy$-plane and let its four sides parallel to $x$ and $y$ axis, respectively. The placement of rectangle $i (i=1,2,\cdots, n)$ in the container can be described by three variables $(x_i, y_i v_i)$, where $x_i, y_i \in \Bbb{N}=\{0,1,2,\cdots\} $ is the coordinate of its bottom-left corner point, $v_i \in \{0,1\}$ denotes its orientation. $v_i = 1$ means it is vertically placed, $v_i= 0$ horizontally. A packing pattern of $n$ rectangles can be described by a vector of $3n$ elements: $\mathcal{X}=(x_1, y_1, v_1, x_2, y_2, v_2,\cdots, x_n, y_n, v_n)$. We give the following definitions.
\newdefinition{mydef}{Definition}
\begin{mydef}[Feasible Packing]
A feasible (or non-overlapping) packing $\mathcal{X}$ satisfies the following two conditions:
\begin{enumerate}[(1)]
\item Each rectangle does not overstep each border of the container.
\item The overlapping area between any two rectangles is zero.
\end{enumerate}
\end{mydef}
\begin{figure}
\begin{center}
\includegraphics[width=1.2in]{./j-over-i.pdf}
\caption{Rectangles 2 and 3 are over rectangle 1}
\label{fig:j-over-i}
\end{center}
\end{figure}
\begin{mydef}[Rectangle $j$ Over Rectangle $i$]
We say rectangle $j$ is over rectangle $i$ (or rectangle $i$ is under rectangle $j$) if and only if there exists a positive number $d$ such that if rectangle $i$ moves upwards by a distance of $d$, then the overlapping area between rectangles $i$ and $j$ is greater than zero. See Fig.\ref{fig:j-over-i}, rectangles 2 and 3 are over rectangle 1, rectangles 4 and 5 are not over rectangle 1. We say rectangle $i$ \textit{ can move upwards freely} if and only if no rectangle is over rectangle $i$.
\end{mydef}
\begin{figure}
\begin{center}
\includegraphics[height=0.9 in]{./j-on-the-right-of-i.pdf}
\caption{Rectangles 2 and 3 are on the right of rectangle 1}
\label{fig:j-on-the-right-of-i}
\end{center}
\end{figure}
\begin{mydef}[Rectangle $j$ On the Right of Rectangle $i$]
We say rectangle $j$ is on the right of rectangle $i$ (or rectangle $i$ is on the left of rectangle $j$) if and only if there exists a positive number $d$ such that if rectangle $i$ moves rightwards by a distance of $d$, then the overlapping area between rectangles $i$ and $j$ is greater than zero. See Fig.\ref{fig:j-on-the-right-of-i}, rectangles 2 and 3 are on the right of rectangle 1, rectangles 4 and 5 are not on the right of rectangle 1. We say rectangle $i$ \textit{can move rightwards freely} if and only if no rectangle is on the right of rectangle $i$.
\end{mydef}
\begin{figure}
\begin{center}
\includegraphics[width=3in]{./bottom-left-corner.pdf}
\caption{Bottom-left stability and bottom-left corners}
\label{fig:blc}
\end{center}
\end{figure}
\begin{mydef}[Bottom-Left Stability]
In a feasible packing, a rectangle is \textit{ bottom-left stable} if and only if it can not move downwards or leftwards without overlapping others \cite{chazelle1983}. A feasible packing is \textit{bottom-left stable} if and only if each rectangle in this packing is bottom-left stable. See Fig.\ref{fig:blc}, rectangles 1, 2, 3, 4 are bottom-left stable; rectangles 5, 6, 7, 8 are not bottom-left stable.
\end{mydef}
\begin{mydef}[Bottom-Left Corner]
A bottom-left corner is an unoccupied area where a suitable rectangle has bottom-left stability. See Fig.\ref{fig:blc}, A, B, C, D, E, F, G are bottom-left corners.
\end{mydef}
\begin{mydef}[Corner Occupying Action]
A corner occupying action is an action that places a rectangle onto a bottom-left corner and makes that rectangle bottom-left stable. Let the placed rectangle be rectangle $i$, then \textit{the rectangles forming the bottom-left corner where rectangle $i$ locates} satisfy the following two conditions: (1) they touch rectangle $i$; (2) they are under or on the left of rectangle $i$.
\end{mydef}
\section{Corner Occupying Theorem}
This section proves the corner occupying theorem. We first present two lemmas.
\newtheorem{mylem}{Lemma}
\newproof{myprf}{Proof}
\begin{mylem}
Any feasible packing can be replaced by another feasible packing where each rectangle has bottom-left stability.
\end{mylem}
\begin{figure}
\begin{center}
\subfigure[Not bottom-left stable]{
\includegraphics[width=1.4in]{./lemma1_a.pdf}}
\subfigure[Bottom-left stable]{
\includegraphics[width=1.4in]{./lemma1_b.pdf}}
\end{center}
\caption{A feasible packing and its equivalent bottom-left stable packing}
\label{fig:lemma1}
\end{figure}
\begin{myprf}
See Fig.\ref{fig:lemma1}, the left packing can be replaced by the right bottom-left stable one. Given a feasible packing $\mathcal{X}_0$, we prove that an equivalent bottom-left stable packing can be found from $\mathcal{X}_0$. Keep the orientation of each rectangle unchanged, suppose that each rectangle can move freely and consider the following function:
\begin{equation}
O = O(x_1, y_1, x_2, y_2, \cdots, x_n, y_n) = \sum_{i=0}^{n-1}{ \sum_{j=i+1}^{n}{O_{ij}} }
\end{equation}
where $O_{ij} (i,j=1,2,\cdots,n)$ is the overlapping area between rectangles $i$ and $j$. $O_{0j} (j=1,2,\cdots, n)$ is the overlapping area between rectangle $j$ and the \textbf{outside} of the container. Let $S_0$ be the set of all zero points of $O$: $S_0 = \{ (x_1, y_1, x_2, y_2, \cdots, x_n, y_n) | O(x_1, y_1, x_2, y_2, \cdots, x_n, y_n) = 0 \}$. Then each point in $S_0$ corresponds to a non-overlapping packing. Because $\mathcal{X}_0$ corresponds to a zero point of $O$, $S_0$ is not empty. And because each rectangle can only be placed at integer coordinate positions, $S_0$ is a finite set.
Then let's consider another function $L$ defined on $S_0$:
\begin{equation}
L = \sum_{i=1}^{N} (x_i + y_i)
\end{equation}
Because $S_0$ is a non-empty and finite set, there exists a point $(x_1^*, y_1^*, x_2^*, y_2^*, \cdots, x_n^*, y_n^*)$ in $S_0$ where $L$ is minimal. Note that $(x_1^*, y_1^*, x_2^*, y_2^*, \cdots, x_n^*, y_n^*)$ corresponds to a feasible packing where each rectangle can not move downwards or leftwards without overlapping others; otherwise, we can find another point in $S_0$ with a smaller $L$, contradicting the fact that $L$ attains its minimum at $(x_1^*, y_1^*, x_2^*, y_2^*, \cdots, x_n^*, y_n^*)$. $\square$
\end{myprf}
Lemma 1 has been explicitly mentioned by \cite{christofides1977, hadji1995, martello2000} as a conjecture and implicitly used by many algorithms for solving the rectangle packing problem.
\begin{mylem}[Escaping Lemma]
In any feasible packing, if we take away the four borders of the container, then there is a rectangle which can move upwards and rightwards freely.
\end{mylem}
\begin{figure}
\begin{center}
\subfigure[]{
\includegraphics[width=1.4in]{./lemma2_1.pdf}}
\subfigure[]{
\includegraphics[width= 1.4in]{./lemma2_2.pdf}}
\caption{Examples for Lemma 2}
\label{fig:lemma2}
\end{center}
\end{figure}
\begin{myprf}
See Fig.\ref{fig:lemma2}(a), the highlighted rectangle can move upwards and rightwards freely. Given a feasible packing with $n$ rectangles, we sort the top-right corner points of the rectangles lexicographically by increasing $<x,y>$ and renumber the rectangles according to this order (See Fig.\ref{fig:lemma2}(b)). We search for the rectangle which can move rightwards and upwards freely as follows. First, we consider the highest numbered rectangle among all $n$ rectangles, i.e., rectangle $n$ (12 in Fig.\ref{fig:lemma2}(b)). Its top-right corner point is the rightmost, thus it can move rightwards freely. If no rectangle is over rectangle $n$, then $n$ is the rectangle we want to find. Otherwise, we consider the highest numbered rectangle among all the rectangles over rectangle $n$. Let it be rectangle $i$ (10 in Fig.\ref{fig:lemma2}(b)). Its top-right corner point is the rightmost among all the rectangles over rectangle $n$. Therefore, it can move rightwards freely. If no rectangle is over rectangle $i$, then $i$ is the rectangle we want to find. Otherwise we consider the highest numbered rectangle among all the rectangles over rectangle $i$ and continue the search as described above.
Because there are only finite ($n$) rectangles, the above search will terminate and we can finally find a rectangle which can move upwards and rightwards freely. $\square$
\end{myprf}
\newtheorem{mythem}{Theorem}
\begin{mythem}
For any feasible, bottom-left stable packing, there exists a numbering of $n$ rectangles such that rectangle $i(i=1,2,\cdots,n)$ locates on a bottom-left corner formed by rectangles $1,2,\cdots, i-1$ and the four borders of the container.
\end{mythem}
\begin{figure}
\begin{center}
\subfigure[Numbering According to Lemma 2]{
\includegraphics[width=1.4in]{./theorem1_1.pdf}}
\subfigure[New Numbering]{
\includegraphics[width= 1.4in]{./theorem1_2.pdf}}
\caption{Examples for Theorem 1}
\label{fig:theorem1}
\end{center}
\end{figure}
\begin{myprf}
According to Lemma 2, there is always a rectangle which can move rightwards and upwards freely in any feasible packing. Consequently, for a feasible packing with $n$ rectangles, we can empty the container by successively taking out a rectangle which can move rightwards and upwards freely. We then number the rectangles according to the order in which rectangles are taken out. Under this numbering, a higher numbered rectangle is not over or on the right of a lower numbered rectangle (See Fig.\ref{fig:theorem1}(a)).
Now we consider a new numbering which is the reversion of the previous numbering, i.e., the number of rectangle $i(i=1,2,\cdots,n)$ becomes $n-i+1$ (See Fig.\ref{fig:theorem1}(b)). Under this new numbering, a lower numbered rectangle is not over or on the right of a higher numbered rectangle. Then, for rectangle $i(i=1,2,\cdots,n)$, the labeled numbers of the rectangles forming the bottom-left corner where rectangle $i$ locates are smaller than $i$.
The new numbering of the rectangles can be taken as an order in which rectangles are placed into the container. Under this order, when we place rectangle $i$ into the container, rectangles $1,2,\cdots, i-1$ are already in the container and rectangles $i+1, i+2, \cdots, n$ are not. Thus, it is shown that, any feasible, bottom-left stable packing can be achieved through a sequence of placement actions, among which the $i$th $(i=1,2,\cdots, n)$ action is to place rectangle $i$ onto a bottom-left corner formed by rectangles $1,2,\cdots, i-1$ and the four borders of the container. That is to say, any feasible, bottom-left stable packing can be achieved through a sequence of corner occupying actions.
$\square$
\end{myprf}
\begin{mythem}[Corner Occupying Theorem]
Arbitrarily given $N$ rectangles and a rectangular container, if it is possible to orthogonally place all the rectangles into the container without overlapping, then we can find a feasible packing through a sequence of corner occupying actions.
\end{mythem}
\begin{myprf}
According to Lemma 1, there exists a feasible packing where each rectangle has bottom-left stability. Then according to Theorem 1, this bottom-left stable packing can be found through a sequence of corner occupying actions. $\square$
\end{myprf}
Note that, in the above proof, the escaping lemma determines the order in which rectangles are placed into the container. And we can find that, when we place rectangle $i$ onto a bottom-left corner in the container, no rectangle in the container is over or on the right of rectangle $i$. Therefore, in practical implementation, in order to reduce the computing time, the corner occupying theorem can be enhanced as : \textit{Arbitrarily given $N$ rectangles and a rectangular container, if it is possible to orthogonally place all the rectangles into the container without overlapping, then we can find a feasible packing by successively placing a rectangle onto a bottom-left corner in the container where the placed rectangle is then under and on the left of no rectangle .}
\section{The Three-dimensional Case}
In this section, we investigate the following problem: can the lemmas and theorems presented in Section 3 be extended to the three-dimensional case? Scientists have done some research for the three-dimensional case\cite{boef2005,martello2000}.
In the three-dimensional case, rectangle $j$ has width $w_j$, height $h_j$ and depth $d_j$, and the container is of width $W$, height $H$ and depth $D$. Similarly, we can extend the notations and definitions presented in Section 2 to the three-dimensional case, and get new definitions like: rectangle $j$ in front of rectangle $i$ (or rectangle $i$ in behind of rectangle $j$), bottom-left-behind corner, bottom-left-behind stability. And then the corner occupying action is to place a rectangle onto a bottom-left-behind corner in the container.
We find that lemma 1 and its proof can be easily extended to the three-dimensional case. However, the escaping lemma is wrong in the three-dimensional case. Fig.\ref{fig:3d} presents a counterexample where no rectangle can move along the positive x-axies, positive y-axies, and positive z-axies freely. The coordinate position of each rectangle in Fig.\ref{fig:3d} is presented in Table \ref{tbl:3d}.
\begin{table*}
\caption{Coordinate position of each rectangle in Fig.\ref{fig:3d} }
\label{tbl:3d}
\begin{center}
\begin{tabular}[htbp]{ccccccc}
\hline
Rectangle & $x$ & $y$ & $z$ & $width$ & $depth$ & $height$ \\
\hline
1 & 2 & 0 & 0 & 1 & 2 & 1 \\
2 & 0 & 1 & 1 & 3 & 1 & 1 \\
3 & 0 & 0 & 2 & 1 & 2 & 1 \\
4 & 1 & 0 & 0 & 1 & 1 & 3 \\
\hline
\end{tabular}
\end{center}
\end{table*}
As shown in Section 3, in the two-dimensional case, the proof of the corner occupying theorem is based on lemma 1 and the escaping lemma. Since the escaping lemma is wrong in the three-dimensional case, we can not directly extend the proof of the corner occupying theorem to the three-dimensional case. However, the incorrectness of escaping lemma in the three-dimensional case might not imply that the corner occupying theorem is incorrect in the three-dimensional case. Therefore, we get an open problem: is the corner occupying theorem correct in the three-dimensional case? Or specifically,
\textit{In the three-dimensional case, if it is possible to orthogonally place $n$ arbitrarily given rectangles into a rectangular container without overlapping, can we achieve a feasible packing by successively placing a rectangle onto a bottom-left-behind corner in the container?}
\begin{figure}
\begin{center}
\includegraphics[width=3in]{./wb3d_2.pdf}
\caption{A counterexample of escaping lemma in the three-dimensional case}
\label{fig:3d}
\end{center}
\end{figure}
|
\section{Amplitude analysis}
\label{sec:DP_analysis}
In Secs.~\ref{sec:kinematics} and \ref{sec:SquareDP} we describe the DP formalism and introduce
the signal parameters that are extracted from
data. In Sec.~\ref{sec:dp_selection} we describe the requirements used to select the signal candidates and
suppress backgrounds. In Sec.~\ref{sec:dp_likelihood} we describe the fit method
and the approach used to account for experimental effects such
as resolution. In Sec.~\ref{sec:fitresults} we present the results of the fit,
and finally, in Sec.~\ref{sec:dp_systematics} we discuss systematic
uncertainties in the results.
\subsection{Decay amplitudes}
\label{sec:kinematics}
The $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ decay contains three identical particles in the final state and therefore the amplitude needs to be symmetrized.
We consider the decay of a spin-zero $\ensuremath{B^0}\xspace$
into three daughters, $\ensuremath{K^0_S}(1)$, $\ensuremath{K^0_S}(2)$, and $\ensuremath{K^0_S}(3)$, with four-momenta $p_1$, $p_2$, and $p_3$. The decay amplitude is given by~\cite{Cheng:2005ug}
\begin{eqnarray}
\label{eq:paper_soni_2}
{\cal A}[\ensuremath{B^0}\xspace&\ensuremath{\rightarrow}\xspace&\ensuremath{K^0_S}(1)\ensuremath{K^0_S}(2)\ensuremath{K^0_S}(3)] \\
\nonumber &=& \left(\frac{1}{2}\right)^{3/2}\Big\{{\cal A}_1[\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\Kbar^0}\xspace(1)\ensuremath{K^0}\xspace(2)\ensuremath{K^0}\xspace(3)] \\
\nonumber &\ & \ \ \ \ \ \ \ \ \ + \ {\cal A}_2[\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\Kbar^0}\xspace(2)\ensuremath{K^0}\xspace(3)\ensuremath{K^0}\xspace(1)] \\
\nonumber &\ & \ \ \ \ \ \ \ \ \ + \ {\cal A}_3[\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\Kbar^0}\xspace(3)\ensuremath{K^0}\xspace(1)\ensuremath{K^0}\xspace(2)]\Big\},
\end{eqnarray}
which takes into account the three permitted paths from the initial state to the final state. For instance for the \ensuremath{B^0}\xspace\ decay this consists of an intermediate state $K^0 K^0 \kern 0.2em\overline{\kern -0.2em K}{}\xspace^0$.
Since the labeling of the three identical particles is arbitrary, we classify the final-state particles according to the square of the invariant mass, $s_{ij}$, defined as
\begin{equation}
\label{eq:minmaxVariables}
s_{ij} \;=\;s_{ji} \;=\; m_{\ensuremath{K^0_S}(i)\ensuremath{K^0_S}(j)}^2 \;=\; (p_i + p_j)^2,
\end{equation}
where $i$ and $j$ are the \ensuremath{K^0_S}\ indices.
We use as independent (Mandelstam) variables the minimum and the maximum of the squared masses $\smin$ and $\smax$:
\begin{eqnarray}
\label{eq:dalitzVariables}
\smin \;&=&\; \min(s_{12},s_{23},s_{13}),\\ \nonumber
\smax \;&=&\; \max(s_{12},s_{23},s_{13}).
\end{eqnarray}
The third (median) invariant squared mass $\smed$ can be obtained from energy and momentum conservation:
\begin{equation}
\label{eq:magicSum}
\smed \;=\; m_{\ensuremath{B^0}\xspace}^2 + 3m_{\ensuremath{K^0_S}}^2
- \smin - \smax.
\end{equation}
The differential $\ensuremath{B}\xspace$ meson decay width with respect to the
variables defined in Eq.~\eqref{eq:dalitzVariables} (i.e.,~the
DP variables) reads
\begin{equation}
\label{eq:partialWidth}
d\Gamma(\ensuremath{B}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}) \;=\;
\frac{1}{(2\pi)^3}\frac{|{\cal A}|^2}{32 m_{\ensuremath{B^0}\xspace}^3}\,d\smin d\smax,
\end{equation}
where ${\cal A}$ is the Lorentz-invariant amplitude
of the three-body decay.
This amplitude analysis does not take into account any flavor tagging or time dependence, thus it is \ensuremath{C\!P}\xspace\ averaged and time integrated. The term $\left|{\cal A}\right|^2$ is therefore simply the average of squares of the contributions ${\cal A}[\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}]$ and ${\cal A}[\ensuremath{\Bbar^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$].
The choice of the variables $\smin$ and $\smax$ gives a uniquely defined coordinate in the symmetrized DP. Therefore only one sixth of the DP is populated, i.e., the event density is six times larger compared to an amplitude analysis involving three distinct particles.
We describe the distribution of signal events
in the DP using an isobar approximation,
which models the total amplitude as
resulting from a coherent sum of amplitudes from the $N$ individual decay channels
of the $B$ meson, either into an intermediate resonance and a bachelor particle or in a nonresonant manner:
\begin{eqnarray}
\label{eq:isobar}
{\cal A}(\smin,\smax)
& = & \sum_{j=1}^{N} c_j F_j(\smin,\smax).
\end{eqnarray}
Here $F_j$ (described in detail below) are DP-dependent amplitudes containing the decay dynamics
and $c_j$ are complex coefficients describing the relative
magnitudes and phases of the different decay channels.
This description, which contains a single complex number $c_j$ for each decay channel regardless of the \ensuremath{B}\xspace flavor (\ensuremath{B^0}\xspace\ or \ensuremath{\Bbar^0}\xspace), reflects the assumptions of no direct \ensuremath{C\!P}\xspace\ violation and of
a common weak phase for all the decay channels.
With this description we cannot extract any weak phase information; this would require using per-\ensuremath{B}\xspace\-flavor complex amplitudes.
The factor $F_j$ contains strong dynamics only, and thus does not change under \ensuremath{C\!P}\xspace\ conjugation.
Intermediate resonances decay to $\ensuremath{K^0}\xspace\!\ensuremath{\Kbar^0}\xspace$. In terms of the isobar approximation, the amplitude in Eq.~\eqref{eq:paper_soni_2} for a resonant state $j$ becomes
\begin{eqnarray}
\label{eq:ResKzKzb}
{\cal A}[\ensuremath{B^0}\xspace&\ensuremath{\rightarrow}\xspace&\ensuremath{K^0_S}(1)\ensuremath{K^0_S}(2)\ensuremath{K^0_S}(3)] \\
\nonumber &\propto& c_j\left[F_j(s_{12},s_{13})+F_j(s_{12},s_{23})+F_j(s_{13},s_{23}) \right].
\end{eqnarray}
This reflects the fact that it is impossible to associate a given \ensuremath{K^0_S}\ to a flavor eigenstate \ensuremath{K^0}\xspace\ or \ensuremath{\Kbar^0}\xspace. In practice, this sum of three $F_j$ terms, corresponding to an even-spin resonance, is implicitly taken into account by the description in terms of $\smin$ and $\smax$.
The $F_j$ terms are represented
by the product of the invariant mass and angular distributions; i.e.,
\begin{equation}
\label{eq:ResDynEqn}
F_j(\smin,\smax, L) = R_j(m) X_L(|\vec{p}\,^{\star}|\,r') X_L(|\vec{q}\,|\,r) T_j(L,\vec{p},\vec{q}\,),
\end{equation}
where
\begin{enumerate}[(i)]
\item $m$ is the invariant mass of the decay products of the resonance;
\item $R_j(m)$ is the resonance mass term or ``line shape'', e.g.,~relativistic
Breit--Wigner (RBW);
\item $L$ is the orbital angular momentum between the resonance and the
bachelor particle;
\item $\vec{p}\,^{\star}$ is the momentum of the bachelor particle
evaluated in the rest frame of the $B$;
\item $\vec{p}$ and $\vec{q}$ are the momenta of the bachelor particle and
one of the resonance daughters, respectively, both evaluated in the
rest frame of the resonance;
\item $X_L(|\vec{p}\,^{\star}|\,r')$ and $X_L(|\vec{q}\,|\,r)$ are Blatt--Weisskopf barrier factors~\cite{blatt-weisskopf}
with barrier radii of $r$ and $r'$, and
\item $T_j(L,\vec{p},\vec{q})$ is the angular distribution:
\begin{eqnarray}
L=0 &:& T_j = 1,\\
L=2 &:& T_j = \frac{8}{3} \left[3(\vec{p}\cdot\vec{q}\,)^2 - (|\vec{p}\,||\vec{q}\,|)^2\right].
\end{eqnarray}
\end{enumerate}
The Blatt--Weisskopf barrier factor is unity for all the zero-spin resonances. In our analysis it is relevant only for the $\fVII$.
Since for this resonance $r$ and $r'$ are not measured, we take them both to be $1.5\ensuremath{\mathrm{\,Ge\kern -0.1em V^{-1}}}\xspace$ and vary by $\pm 0.5\ensuremath{\mathrm{\,Ge\kern -0.1em V^{-1}}}\xspace$ to estimate the systematic uncertainty.
The helicity angle of a resonance is defined as the angle between $\vec{p}$ and $\vec{q}$.
Explicitly, the helicity angle $\theta$
for a given resonance
is defined between the momenta of the bachelor particle and one of the daughters of the resonance in the resonance rest frame.
Because of the identical final-state particles this definition is ambiguous,
but the ambiguity disappears because of the description of the DP in terms of \smin\ and \smax.
There are three possible invariant-mass combinations: $\smin$, $\smed$, and $\smax$. We denote the corresponding helicity angles as $\theta_{\min}$, $\theta_{\rm med}$, and $\theta_{\max}$.
The three angles are defined between $0$ and $\pi/2$.
As the present study is the first amplitude analysis of this decay, we use the method outlined in Sec.~\ref{sec:signal_model} to determine the contributing intermediate states. The components of the nominal signal model are summarized in Table~\ref{tab:model}.
For most resonances in this analysis the $R_j$ are taken to be RBW~\cite{Amsler:2008zz} line shapes:
\begin{equation}
R_j(m) = \frac{1}{(m^2_0 - m^2) - i m_0 \Gamma(m)},
\label{eqn:BreitWigner}
\end{equation}
where $m_0$ is the nominal mass of the resonance and $\Gamma(m)$ is the
mass-dependent width.
In the general case of a spin-$J$ resonance, the latter can be expressed as
\begin{equation}
\Gamma(m) = \Gamma_0 \left( \frac{q}{q_0}\right)^{2J+1}
\left(\frac{m_0}{m}\right) \frac{X^2_J(|\vec{q}\,|\,r)}{X^2_J(|\vec{q_0}\,|\,r)}.
\label{eqn:resWidth}
\end{equation}
The symbol $\Gamma_0$ denotes the nominal width of the resonance.
The values of $m_0$ and $\Gamma_0$ are listed in Table~\ref{tab:model}.
The symbol $q_0$ denotes the value of $q$ when $m = m_0$.
For the \fI\ line shape the Flatt\'e form~\cite{Flatte} is used.
In this case the mass-dependent width is given by the sum
of the widths in the $\pi\pi$ and $KK$ systems:
\begin{equation}
\Gamma(m) = \Gamma_{\pi\pi}(m) + \Gamma_{KK}(m),
\label{eqn:FlatteW1}
\end{equation}
where
\begin{eqnarray}
\Gamma_{\pi\pi}(m) &=&
g_{\pi} \Bigg(\frac{1}{3}\sqrt{1 - 4m_{\ensuremath{\pi^0}\xspace}^2/m^2} + \\ \nonumber
&& \phantom{g_{\pi} \Bigg(\frac{1}{3}} \frac{2}{3}\sqrt{1 - 4m_{\ensuremath{\pi^\pm}\xspace}^2/m^2}\Bigg)\,,\\
\Gamma_{KK}(m) &=&
g_{K} \Bigg(\frac{1}{2}\sqrt{1 - 4m_{\ensuremath{K^\pm}\xspace}^2/m^2} + \\ \nonumber
&& \phantom{g_{K} \Bigg(\frac{1}{2}} \frac{1}{2}\sqrt{1 - 4m_{\ensuremath{K^0}\xspace}^2/m^2}\Bigg)\,.
\label{eqn:FlatteW2}
\end{eqnarray}
The fractional coefficients arise from isospin conservation and $g_{\pi}$
and $g_{K}$ are coupling constants for which the values are given in Table~\ref{tab:model}.
The nonresonant (NR) component is modeled using an exponential function:
\begin{equation}
R_{\rm NR}(m) = e^{\alpha m^2}.
\label{eqn:NREXP}
\end{equation}
As in the resonant case, here $m$ is the invariant mass of the relevant \ensuremath{K^0_S}\KS\ pair.
The parameter $\alpha$ is taken from the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ $\ensuremath{\Bu}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{K^+}\xspace\ensuremath{K^-}\xspace\ensuremath{K^+}\xspace$ analysis~\cite{Aubert:2007sd,babarKKKs} and is given in Table~\ref{tab:model}. This value was found to be compatible with the one resulting from varying $\alpha$ in the maximum-likelihood fit in the present analysis.
There is no satisfactory theoretical description of the NR component; it has to be determined empirically. The exponential function of Eq.~\eqref{eqn:NREXP} was used by other amplitude analyses of \ensuremath{B}\xspace-meson decays to three kaons~\cite{babarKKK, Aubert:2007sd, babarKKKs, belleKKK, belleKKKs}. Adopting the same parametrization for the NR term allows the comparison of results for other components.
\begin{table*}[htbp]
\begin{center}
\caption{Parameters of the DP model used in the fit.
The Blatt--Weisskopf barrier parameters ($r$ and $r'$) of the $\fVII$, which have not been measured, are varied by $\pm 0.5\ensuremath{\mathrm{\,Ge\kern -0.1em V^{-1}}}\xspace$ for the model uncertainty.
\label{tab:model}}
\begin{tabular}{cccc}
\hline\hline
Resonance & Parameters & Line shape & Reference \\ \hline\\[-9pt]
$\fI$ & $m_0=(965 \pm 10)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & Flatt\'e & \cite{valFlatte} \\
& $g_{\pi}=(165 \pm 18)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & Eq.~\eqref{eqn:FlatteW1} & \\
& $g_{K}=(695 \pm 93)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & & \\ \hline\\[-9pt]
$\fV$ & $m_0=(1724 \pm 7)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & RBW & \cite{Amsler:2008zz} \\
& $\Gamma_0=(137 \pm 8)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & Eq.~\eqref{eqn:BreitWigner} & \\ \hline\\[-9pt]
$\fVII$ & $m_0=(2011\,^{+60}_{-80})\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$& RBW & \cite{Amsler:2008zz} \\
& $\Gamma_0=(202 \pm 60)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & Eq.~\eqref{eqn:BreitWigner} & \\ \\[-9pt]
& $r=r'=1.5 \ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace^{-1}$ & & \\ \hline\\[-9pt]
NR decays \hspace*{5 mm}& $\alpha=(-0.14 \pm 0.02)\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace^{-2} c^4$ \hspace*{5 mm}& Exponential NR & \cite{babarKKKs} \\
& & Eq.~\eqref{eqn:NREXP} & \\ \hline\\[-9pt]
$\ensuremath{\chi_{c0}}\xspace$ & $m_0=(3414.75 \pm 0.31)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & RBW & \cite{Amsler:2008zz} \\
& $\Gamma_0=(10.2 \pm 0.7)\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ & Eq.~\eqref{eqn:BreitWigner} & \\
\hline \hline
\end{tabular}
\end{center}
\end{table*}
\subsection{The square Dalitz plot}
\label{sec:SquareDP}
We use two-dimensional histograms to describe the phase-space dependent reconstruction efficiency and to model the background over the DP. When the phase-space boundaries of the DP do not coincide with the histogram bin boundaries this may introduce biases. We therefore define $\hmin$ and $\hmax$ as
$\cos \theta_{\min}$ and $\cos \theta_{\max}$, respectively, and apply the transformation
\begin{equation}
\label{eq:SqDalitzTrans1}
\left(\smin, \smax\right) \longrightarrow \left(\hmin, \hmax\right).
\end{equation}
The $\left(\hmin, \hmax\right)$ plane is referred to as the square Dalitz plot (SDP), where both $\hmin$ and $\hmax$ range between $0$ and $1$
due to the convention adopted for the helicity angles (see Fig.~\ref{fig:SquareDalitz}).
Explicitly, the transformation is
\begin{eqnarray}
\hmin &=&
\frac{\smin(\smax-\smed)}{\sqrt{\smin^2-4m_{\KS}^2\smin}}\times \\ \nonumber
&& \frac{1}{\sqrt{(m_{\ensuremath{B^0}\xspace}^2-m_{\KS}^2-\smin)^2-4m_{\KS}^2\smin}}, \\
\hmax &=&
\frac{\smax(\smed-\smin)}{\sqrt{\smax^2-4m_{\KS}^2\smax}}\times \\ \nonumber
&& \frac{1}{\sqrt{(m_{\ensuremath{B^0}\xspace}^2-m_{\KS}^2-\smax)^2-4m_{\KS}^2\smax}}, \nonumber
\label{eq:SqDalitzTransExplicit}
\end{eqnarray}
where the numerators may easily be expressed in terms of \smin\ and \smax\ using Eq.~\eqref{eq:magicSum}.
The differential surface elements of the DP and the SDP are related by
\begin{equation}
\label{eq:SqDalitzTrans2}
d \smin \,d \smax = |\det J|\, d \hmin \, d \hmax,
\end{equation}
where $J=J\left(\hmin, \hmax\right)$ is the appropriate Jacobian matrix.
The backward transformations $\smin\left(\hmin, \hmax\right)$ and $\smax\left(\hmin, \hmax\right)$, and therefore the Jacobian $|\det J|$, cannot be found analytically; they are obtained numerically.
The variables \hmin\ and \hmax\ as a function of the invariant masses are shown in Fig.~\ref{fig:SquareDalitz} together with the Jacobian.
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=8.0cm,keepaspectratio]{figures/IsoHelLines.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/JacPlot.eps}
\caption{(color online).
Lines of constant helicity angle in the Dalitz plot of \smin\ versus \smax\ (left), and
the magnitude of the Jacobian (gray scale on the right) mapping (\smin, \smax) to (\hmin, \hmax). For the latter see Eq.~\eqref{eq:SqDalitzTrans2}.}
\label{fig:SquareDalitz}
\end{center}
\end{figure*}
\subsection{Event selection and backgrounds}
\label{sec:dp_selection}
We reconstruct $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ candidates from three $\ensuremath{K^0_S}\ensuremath{\rightarrow}\xspace\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$ candidates that form a good quality vertex; i.e., the fit of the \ensuremath{B^0}\xspace\ vertex is required to converge
and the $\chi^2$ probability of each \ensuremath{K^0_S}\ vertex fit has to be greater than $10^{-6}$.
Each $\ensuremath{K^0_S}$ candidate must have $\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$ invariant mass within $12.1\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ of the nominal $\ensuremath{K^0}\xspace$ mass~\cite{Amsler:2008zz}, and decay length with respect to the $B$ vertex between $0.22$ and $45\ensuremath{{\rm \,cm}}\xspace$. The last criterion ensures that the decay vertices of the \ensuremath{B^0}\xspace and the \ensuremath{K^0_S}\ are well separated.
In addition, combinatorial background is suppressed by selecting events for which the angle between
the momentum vector of each \ensuremath{K^0_S}\ candidate
and the vector connecting the beamspot and the $\ensuremath{K^0_S}$ vertex is smaller than
$0.0185$~radians.
We ensure a good $B$ vertex fit quality by requiring that the charged pions of at least one of the $\ensuremath{K^0_S}$ candidates have hits in the two inner layers of the vertex tracker.
A $B$ meson candidate is characterized kinematically by the energy-substituted mass $\mbox{$m_{\rm ES}$}\xspace\equiv\sqrt{(s/2+\vec {p}_i\cdot\vec{p}_B)^2/E_i^2-p_B^2}$ and the energy difference $\Delta E \equiv E_B^*-\mbox{$\frac{1}{2}$}\sqrt{s}$, where $(E_B,\vec{p}_B)$ and $(E_i,\vec {p}_i)$ are the four-vectors in the laboratory frame of the $B$-candidate and the initial electron-positron system,
respectively, and $p_B$ is the magnitude of $\vec{p}_B$. The asterisk denotes the \Y4S\ frame, and $s$ is the square of the invariant mass of the electron-positron system. We require $5.27 < \mbox{$m_{\rm ES}$}\xspace <5.29\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ and $|\Delta E|<0.1\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$. Following the calculation of these kinematic variables, each of the $B$ candidates is refitted with its mass constrained to the world average value of the $B$ meson mass~\cite{Amsler:2008zz} in order to improve the DP position resolution, and ensure that Eq.~\eqref{eq:magicSum} holds. The sideband used for background studies is in the range $5.20 < \mbox{$m_{\rm ES}$}\xspace <5.27\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ and $|\Delta E|<0.1\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$.
Backgrounds arise primarily from random combinations in continuum $e^+e^- \ensuremath{\rightarrow}\xspace q \bar{q}$ events ($q=d,u,s,c$). To enhance discrimination between signal and continuum background, we use a neural network (NN)~\cite{Gay:1995sm} to combine four discriminating variables: the angles with respect to the beam axis of the $B$ momentum and $B$ thrust axis in the \Y4S\ frame, and the zeroth- and second-order monomials $L_{0,2}$ of the energy flow about the $B$ thrust axis. The monomials are defined by $ L_n = \sum_i p_i\times\left|\cos\theta_i\right|^n$, where $\theta_i$ is the angle with respect to the $B$ thrust axis of track or neutral cluster $i$ and $p_i$ is the magnitude of its momentum. The sum excludes the $B$ candidate and all quantities are calculated in the \Y4S\ frame. The NN is trained with off-resonance data, sideband data and simulated signal events that pass the selection criteria.
Approximately $0.5\%$ of events passing the full selection have more than one candidate. When this occurs, we select the candidate for which the error-weighted average of the masses of the \ensuremath{K^0_S}\ candidates is closest to the world average \ensuremath{K^0_S}\ mass~\cite{Amsler:2008zz}.
With the above selection criteria, we obtain a signal reconstruction efficiency of $6.6\%$ that has been determined
from a signal Monte Carlo (MC) sample generated using
the same DP model and parameters as obtained from the data fit results.
We estimate from this MC that $1.4\%$ of the selected signal events are misreconstructed, and assign a systematic uncertainty (see Sec.~\ref{sec:dp_systematics}).
We use MC events to study the background from other $B$ decays ($B$ background). We expect fewer than 6 such events in our data sample. As these events are wrongly reconstructed, the $\mbox{$m_{\rm ES}$}\xspace$ and $\Delta E$ distributions are continuumlike and as a result the events are mostly absorbed in the continuum background category. We assign a systematic uncertainty for $B$ background contamination in the signal.
\subsection{Results}
\label{sec:fitresults}
The maximum-likelihood fit of $505$ candidates results in a $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ event yield of $200\pm 15$ and a continuum yield of $305\pm 18$, where the uncertainties are statistical only.
The symmetrized and square Dalitz plots of a signal DP-model MC sample generated with the result of the fit to data are shown in Fig.~\ref{fig:MCsignal}.
Figure~\ref{fig:sPlots} shows plots of $\Delta E$, $\mbox{$m_{\rm ES}$}\xspace$, and the NN for isolated signal and continuum background events obtained by the {\ensuremath{\hbox{$_s$}{\cal P}lots}}~\cite{Pivk:2004ty} technique.
Figure~\ref{fig:projPlots} shows projections of the data onto the invariant masses $\smin$ and $\smax$.
\begin{figure*}[htbp]
\includegraphics[width=8.0cm,keepaspectratio]{figures/Dalitz_Plot_Sym_SP10045_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/SqDalitz_Plot_Sym_SP10045_noTitle.eps}
\caption{
Symmetrized (left) and square (right) DP for MC simulated signal events using the amplitudes obtained from the fit to data. The low population in bins along the edge of the symmetrized DP is due to the fact that the phase space boundaries do not coincide with the histogram bin boundaries.}
\label{fig:MCsignal}
\end{figure*}
\begin{figure*}[htbp]
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Signal_mES_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Cont_mES_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Signal_DeltaE_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Cont_DeltaE_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Signal_NN_noTitle.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/sPlot_Cont_NN_noTitle.eps}
\caption{(color online). {\ensuremath{\hbox{$_s$}{\cal P}lots}} (points with error bars) and PDFs (histograms) of the discriminating variables: $\mbox{$m_{\rm ES}$}\xspace$ (top), $\Delta E$ (middle), and NN (bottom), for signal events (left) and continuum events (right).
Below each bin are shown the residuals, normalized in error units. The horizontal dotted and full lines mark the one- and two-standard deviation levels, respectively.
}
\label{fig:sPlots}
\end{figure*}
\begin{figure*}[htbp]
\includegraphics[width=8.0cm,keepaspectratio]{figures/mMin_Projection_Nom.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/mMax_Projection_Nom.eps}
\caption{(color online). Projections onto $\sqrt{\smin}$ (left) and $\sqrt{\smax}$ (right). On-resonance data are shown as points with error bars while the dashed (dotted) histogram represents the signal (continuum) component. The solid-line histogram is the total PDF. Below each bin are shown the residuals, normalized in error units. The horizontal dotted and full lines mark the one- and two-standard deviation levels, respectively.}
\label{fig:projPlots}
\end{figure*}
When the fit is repeated with initial parameter values randomly chosen within wide ranges above and below the nominal values for the magnitudes and within the $[-\pi,\,\pi]$ interval for the phases, we observe convergence towards two solutions with minimum values of the negative log likelihood function $-2\ln{\cal L}$ separated by $3.25$ units. In the following, we refer to them as
Solution 1 (the global minimum) and Solution 2 (a local minimum). No other local minima were found.
In the fit, we measure directly the relative magnitudes and phases of the different components of the signal model. The magnitude and phase of the NR amplitude are fixed to $1$ and $0$, respectively, as a reference.
In Fig.~\ref{fig:ScanMagf0} we show likelihood scans of the isobar magnitudes and phases of all the resonances, where both solutions can be noticed. Each of these scans is obtained by fixing the corresponding isobar parameter at several consecutive values, for each of which
the fit to the data is repeated.
The measured relative amplitudes $c_{\mu}$
are used to extract the fit fraction defined as
\begin{equation}
\label{eq:ff}
FF(k) = \frac{\sum_{\mu=3k -2}^{3k}\sum_{\nu=3k -2}^{3k}{c_{\mu}c^*_{\nu}\langle F_{\mu}F^*_{\nu}
\rangle}}{\sum_{\mu \nu}{c_{\mu}c^*_{\nu}\langle F_{\mu}F^*_{\nu}
\rangle}},
\end{equation}
where $k$, which varies from 1 to 5, represents an intermediate state.
Each fit fraction is a sum of three identical contributions, one for each pair of \ensuremath{K^0_S}. The indices $\mu$ and $\nu$ run from 1 to 15, as each of the five resonances contributes to three pairs of \ensuremath{K^0_S} , which correspond to the three terms ($3k-2$, $3k-1$, and $3k$) in each sum in the numerator of Eq.~\eqref{eq:ff}.
The dynamical amplitudes $F$ are defined in Sec.~\ref{sec:kinematics} and the terms
\begin{equation}
\langle F_{\mu}F^*_{\nu} \rangle = \int\int{F_{\mu}F^*_{\nu}d \smin d \smax}
\end{equation}
are obtained by integration over the DP.
The total fit fraction is defined as the algebraic sum of all fit fractions. This quantity is not
necessarily unity due to the potential presence of net constructive or destructive interference.
In order to estimate the statistical significance of each resonance, we
evaluate the difference $\Delta \ln {\cal L}$ between the log-likelihood of the nominal fit and that of a fit where the magnitude of the amplitude of the resonance is set to $0$ (this difference can be directly read from the likelihood scans as a function of magnitudes in Fig.~\ref{fig:ScanMagf0}). In this case the phase of the resonance becomes meaningless, and we therefore account for two degrees of freedom removed from the fit.
The value $2 \Delta \ln {\cal L}$ is used to evaluate the $p$-value for $2$ degrees of freedom;
we determine the equivalent one-dimensional significance from this $p$-value.
The results for the phase and the fit fraction are given in Table~\ref{tab:Q2B} for the
two solutions; the change in likelihood when the amplitude of the resonance is set to $0$ and the resulting statistical significance of each resonance is given for Solution 1.
As the fit fractions are not parameters of the PDF itself, their statistical errors are obtained from the $68.3\%$ coverage intervals of the fit-fraction distributions obtained from a large number of pseudoexperiments generated with the corresponding solution (1 or 2). As observed in other three-kaon modes~\cite{babarKKK,belleKKK,Aubert:2007sd,babarKKKs,belleKKKs}, the total FF significantly exceeds unity.
\begin{table}
\caption{
Summary of measurements of the quasi-two-body parameters. The quoted uncertainties are statistical only.
The change in the log-likelihood ($-2 \Delta \ln{\cal L}$) corresponds to the case where the magnitude of the amplitude of the resonance is set to 0. This number is used for the estimation of the statistical significance of each resonance.
}
\renewcommand{\arraystretch}{1.5}
\begin{tabular}{llcc}
\hline \hline
Mode & Parameter & Solution 1 & Solution 2\\ \hline
$f_0(980)\ensuremath{K^0_S}$ & FF & $0.44\,^{+0.20}_{-0.19}$ & $1.03\,^{+0.22}_{-0.17}$ \\
& Phase [rad] & $0.09 \pm 0.16$ & $1.26 \pm 0.17$ \\
& $-2 \Delta \ln{\cal L}$ & $11.7$ & - \\
& Significance [$\sigma$] & $3.0$ & - \\ \hline
$f_0(1710)\ensuremath{K^0_S}$& FF & $0.07\,^{+0.07}_{-0.03} $ & $0.09\,^{+0.05}_{-0.02}$ \\
& Phase [rad] & $1.11 \pm 0.23$ & $0.36 \pm 0.20$ \\
& $-2 \Delta \ln{\cal L}$ & $14.2$ & - \\
& Significance [$\sigma$] & $3.3$ & - \\ \hline
$f_2(2010)\ensuremath{K^0_S}$& FF & $0.09\,^{+0.03}_{-0.03}$ & $0.10 \pm 0.02$ \\
& Phase [rad] & $2.50 \pm 0.20$ & $1.58 \pm 0.22$ \\
& $-2 \Delta \ln{\cal L}$ & $14.0$ & - \\
& Significance [$\sigma$] & $3.3$ & - \\ \hline
NR & FF & $2.16\,^{+0.36}_{-0.37}$ & $1.37\,^{+0.26}_{-0.21}$ \\
& Phase [rad] & $0.0$ & $0.0$ \\
& $-2 \Delta \ln{\cal L}$ & $68.1$ & - \\
& Significance [$\sigma$] & $8.0$ & - \\ \hline
$\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0_S}$& FF & $0.07\,^{+0.04}_{-0.02}$ & $ 0.07 \pm 0.02$ \\
& Phase [rad] & $0.63 \pm 0.47$ & $-0.24 \pm 0.52$ \\
& $-2 \Delta \ln{\cal L}$ & $18.5$ & - \\
& Significance [$\sigma$] & $3.9$ & - \\ \hline
& Total FF & $2.84\,^{+0.71}_{-0.66}$ & $2.66\,^{+0.35}_{-0.27}$ \\ \hline \hline
\end {tabular}
\renewcommand{\arraystretch}{1.0}
\label{tab:Q2B}
\end{table}
In Table~\ref{tab:Q2B} it can be seen that the two solutions differ mostly in the fraction assigned to the NR and the $\fI$ components.
Solution 1 corresponds to a small FF of the $\fI$ and a large value for the NR, and Solution 2 has a large $\fI$ fraction and a smaller NR fraction. Other three-kaon modes~\cite{babarKKK,belleKKK,Aubert:2007sd,babarKKKs,belleKKKs} favor the behavior of Solution 1.
\begin{figure*}[htbp]
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Magnitude_f0_980_Nice_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Phase_f0_980_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Magnitude_f0_1710_Nice_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Phase_f0_1710_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Magnitude_f0_2010_Nice_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Phase_f2_2010_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Magnitude_f0_chi_c0_Nice_noTitle.eps}
\includegraphics[width=7.64cm,keepaspectratio]{figures/IsobarScan_Phase_chi_c0_noTitle.eps}
\caption{
(color online). One-dimensional scans of $-2 \Delta \ln{\cal L}$ as a function of magnitudes (left) and phases (right) of the resonances $\fI$, $\fV$, $\fVII$, and $\ensuremath{\chi_{c0}}\xspace$ (top to bottom). The horizontal dashed lines mark the one- and two-standard deviation levels.}
\label{fig:ScanMagf0}
\end{figure*}
Generalizing Eq.~\eqref{eq:ff}, we obtain the interference fractions among the intermediate decay modes $k$ and $j$
\begin{equation}
FF(k,j) = \frac{\sum_{\mu=3k -2}^{3k}\sum_{\nu=3j-2}^{3j}{c_{\mu}c^*_{\nu}\langle F_{\mu}F^*_{\nu}
\rangle}}{\sum_{\mu \nu}{c_{\mu}c^*_{\nu}\langle F_{\mu}F^*_{\nu}
\rangle}},
\end{equation}
which are given in Table~\ref{tab:ff_interference} for Solution 1. Unlike the total FF defined above, the elements of this matrix sum to unity. The large destructive interference between the $\fI\ensuremath{K^0_S}$ and the NR components appears clearly in the table.
This is possible due to the large overlap in phase space between the exponential NR term and the broad tail of the $\fI$ resonance above the $KK$ threshold.
\begin{table*}[htbp]
\centering
\caption{The interference fractions $FF(k,j)$ among the intermediate decay amplitudes for Solution 1. Note that the diagonal elements are those defined in Eq.~\eqref{eq:ff} and detailed in Table~\ref{tab:Q2B}. The lower diagonal elements are omitted since the matrix is symmetric.}
\begin{tabular}{l|ccccc}
\hline\hline
& $\fI\ensuremath{K^0_S}$ & $\fV\ensuremath{K^0_S}$ & $\fVII\ensuremath{K^0_S}$ & NR & $\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0_S}$ \\ \hline
$\fI\ensuremath{K^0_S}$ & $0.44$ & $0.07$ & $-0.02$ & $-0.80$ &$\ \ 0.01$ \\
$\fV\ensuremath{K^0_S}$ & & $0.07$ & $-0.01$ & $-0.17$ & $-0.0003$ \\
$\fVII\ensuremath{K^0_S}$ & & & $\ \ 0.09$ & $\ \ 0.02$&$\ \ 0.0002$ \\
NR & & & & $\ \ 2.16$& $-0.02$ \\
$\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0_S}$ & & & & &$\ \ 0.07$ \\
\hline\hline
\end{tabular}
\label{tab:ff_interference}
\end{table*}
Using the relative fit fractions, we calculate the branching fraction $\mathcal B$ for the intermediate mode $k$ as
\begin{equation}
\label{eq:BF_definition}
FF(k)\times \mathcal{B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S} ),
\end{equation}
where $\mathcal{B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S} )$ is the total inclusive branching fraction:
\begin{equation}
\label{total_inclusive_BF}
\mathcal{B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S})=\frac{N_{\rm sig}}{\bar{\varepsilon} N_{B\bar{B}}}.
\end{equation}
We estimate the average efficiency $\bar{\varepsilon}=6.6\%$ using a fully reconstructed DP-model MC sample generated with the parameters found in data.
The results of the branching fraction measurements are shown in Table~\ref{tab:BR}. As a cross check we attempt to compare our measured branching fractions to results from other measurements; however, many of the branching fractions for the decay into kaons of the resonances included in our model are not (or are only poorly) measured (marked as ``seen" in Ref.~\cite{Amsler:2008zz}). An exception is the charmonium state $\ensuremath{\chi_{c0}}\xspace$, for which the measured value is ${\cal B}(\ensuremath{\chi_{c0}}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS)=(3.16 \pm 0.18)\times 10^{-3}$~\cite{Amsler:2008zz}. We can then use the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ measurement of ${\cal B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0}\xspace)=(142\,^{+55}_{-44} \pm 8 \pm 16 \pm 12)\times 10^{-6}$~\cite{Aubert:2009me} to calculate ${\cal B}[\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\chi_{c0}}\xspace(\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS)\ensuremath{K^0_S}]=\frac{1}{2}\,{\cal B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0}\xspace)\times{\cal B}(\ensuremath{\chi_{c0}}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS)=(0.224 \pm 0.078)\times 10^{-6}$, which is consistent with our measured branching fraction, given in Table~\ref{tab:BR}.
An interesting conclusion from this first amplitude analysis of the $\ensuremath{B^0}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ decay mode is that we do not need to include a broad scalar $f_X(1500)$ resonance, as has been done in other measurements~\cite{babarKKK,belleKKK,Aubert:2007sd,babarKKKs,belleKKKs}, to describe the data. The peak in the invariant mass between $1.5$ and $1.6\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ can be described by the interference between the $f_0(1710)$ resonance and the nonresonant component. However minor contributions from the $f_{2}^{'}(1525)$ and $f_{0}(1500)$ resonances to this structure cannot be excluded.
\begin{table}[htbp]
\caption{Summary of measurements of branching fractions (${\cal B}$). The quoted numbers are obtained by multiplying the corresponding fit fraction from Solution 1 by the measured inclusive $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ branching fraction. The first uncertainty is statistical, the second is systematic, and the third represents the signal DP-model dependence.}
\renewcommand{\arraystretch}{1.5}
\begin{tabular}{lc}
\hline \hline
Mode & ${\cal B}$ [$\times 10^{-6}$] \\ \hline
Inclusive $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ & $6.19 \pm 0.48 \pm 0.15 \pm 0.12$ \\ \hline
$f_0(980)\ensuremath{K^0_S}$, $~~f_0(980)\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS$& $2.7\,^{+1.3}_{-1.2} \pm 0.4 \pm 1.2$ \\
$f_0(1710)\ensuremath{K^0_S}$, $~~f_0(1710)\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS$& $0.50\,^{+0.46}_{-0.24} \pm 0.04 \pm 0.10$ \\
$f_2(2010)\ensuremath{K^0_S}$, $~~f_2(2010)\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS$& $0.54\,^{+0.21}_{-0.20} \pm 0.03 \pm 0.52$ \\
${\rm NR}$, $~~\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$& $13.3\,^{+2.2}_{-2.3} \pm 0.6 \pm 2.1$ \\
$\ensuremath{\chi_{c0}}\xspace\ensuremath{K^0_S}$, $~\ensuremath{\chi_{c0}}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS$& $0.46\,^{+0.25}_{-0.17} \pm 0.02 \pm 0.21$ \\ \hline \hline
\end {tabular}
\renewcommand{\arraystretch}{1.0}
\label{tab:BR}
\end{table}
\subsection{The maximum-likelihood fit}
\label{sec:dp_likelihood}
We perform an unbinned extended maximum-likelihood fit to extract the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ event yield, as well as the resonant and nonresonant amplitudes.
The fit for the amplitude analysis uses the variables $\mbox{$m_{\rm ES}$}\xspace$ and $\Delta E$, the NN output, and the SDP variables to discriminate signal from background. The selected on-resonance data sample is assumed to consist of signal and continuum background.
The feed-through from \ensuremath{B}\xspace decays other than the signal is found to be negligible.
Misreconstructed signal events are not considered as a separate event species, but are taken into account as a part of the signal.
The likelihood function ${\cal L}_i$ for event $i$ is the sum
\begin{equation}
\label{eq:PropDenSingle}
{\cal L}_i=\sum_j{N_j{\cal P}^i_j(\mbox{$m_{\rm ES}$}\xspace,\Delta E,{\rm NN},h_{\min},h_{\max})},
\end{equation}
where $j$ stands for the species (signal, continuum background) and $N_j$ is the corresponding yield. Each probability density function (PDF) ${\cal P}^i_j$ is the product of four individual PDFs:
\begin{equation}
\label{eq:PropProd}
{\cal P}^i_j={\cal P}^i_j(\mbox{$m_{\rm ES}$}\xspace) \; {\cal P}^i_j(\Delta E) \; {\cal P}^i_j({\rm NN}) \; {\cal P}^i_j(h_{\min},h_{\max}).
\end{equation}
A study with fully reconstructed MC samples shows that correlations between the PDF variables are small and therefore we neglect them. However, possible small discrepancies in the fit results due to these correlations are accounted for in the systematic uncertainty (see Sec.~\ref{sec:dp_systematics}).
The total likelihood is given by
\begin{equation}
\label{eq:dp_Likelihood}
{\cal L}=\exp(-\sum_j N_j)\prod_i {\cal L}_i.
\end{equation}
\subsubsection{The \mbox{$m_{\rm ES}$}\xspace, \mbox{$\Delta E$}\xspace, and NN PDFs}
\label{sec:dp_likeDiscrim}
The \mbox{$m_{\rm ES}$}\xspace\ and \mbox{$\Delta E$}\xspace\ distributions of signal events are parametrized by an asymmetric Gaussian with power-law tails:
\begin{eqnarray}
\label{eq:cruijff}
& & {\rm Cr}(x;m_0,\sigma_l,\sigma_r,\alpha_l,\alpha_r)=\\
\nonumber & & \exp\left(-\frac{(x-m_0)^2}{2\sigma_i^2+\alpha_i(x-m_0)^2}\right)
\left\{
\begin{array}{ll}
x-m_0<0 : & i=l\\
x-m_0\geq0 : & i=r
\end{array}.
\right.
\end{eqnarray}%
The $m_0$ parameters for both \mbox{$m_{\rm ES}$}\xspace\ and \mbox{$\Delta E$}\xspace\
are free in the fit to data, while the other parameters are fixed to values determined from a fit to MC simulation. For the NN distributions of signal we use a histogram PDF from MC simulation.
For continuum events the \mbox{$m_{\rm ES}$}\xspace and \mbox{$\Delta E$}\xspace\ PDFs are parametrized by an ARGUS shape function~\cite{argus} and a straight line, respectively. The NN PDF is described by a sum of power functions:
\begin{eqnarray}
\label{eq:spexp}
& & E(x;c_1,a,b_0,b_1,b_2,b_3,c_2,c_3)=\\
\nonumber & & \cos^2(c_1) \: [\cos^2(a)\: {\cal N}(b_0,b_1)\: x^{b_0}\: (1-x)^{b_1} \\
\nonumber & & + \sin^2(a)\:{\cal N}(b_2,b_3)\: x^{b_2}\: (1-x)^{b_3}] \\
\nonumber & & + \sin^2(c_1)\: {\cal N}(c_2,c_3)\: x^{c_2}\: (1-x)^{c_3},
\end{eqnarray}
where
$x=\left({\rm NN}-{\rm NN}_{\rm min}\right)/\left({\rm NN}_{\rm max}-{\rm NN}_{\rm min}\right)$
and the ${\cal N}$ are normalization factors, computed analytically using the standard $\Gamma$ function:
\begin{equation}
{\cal N}(\alpha,\beta) = \frac{\Gamma(\beta+2+\alpha)}{\Gamma(\alpha+1)\Gamma(\beta+1)}.
\end{equation}
The parameters for all the continuum PDFs are determined by a fit to sideband data and then fixed for the fit in the signal region.
\subsubsection{Dalitz-plot PDFs}
\label{sec:dp_likeDalitz}
The SDP PDF for continuum background is a histogram obtained from \mbox{$m_{\rm ES}$}\xspace sideband on-resonance events.
The SDP signal PDFs require as input the DP-dependent selection efficiency, $\varepsilon=\varepsilon(\hmin,\hmax)$, that is described by a histogram and is taken from MC simulation.
For each event we define the SDP signal PDF:
\begin{equation}
\label{eq:sigSDPPDF}
{\cal P}^i_{\rm sig} (\hmin, \hmax)\propto\varepsilon(\hmin, \hmax)\:\left|{\cal A}(\hmin,\hmax)\right|^2.
\end{equation}
The normalization of the PDF is implemented by numerical integration.
To describe the experimental resolution in the SDP variables, we use an ensemble of two-dimensional histograms that represents the probability to reconstruct at the coordinate (\hmin', \hmax') an event that has the true coordinate (\hmin, \hmax). These histograms are taken from MC simulation and are convolved with the signal PDF.
\subsubsection{Determination of the signal Dalitz-plot model}
\label{sec:signal_model}
Using on-resonance data, we determine a nominal signal DP model by making likelihood scans with various combinations of isobars.
We start from a baseline model that includes $\fI$, $\ensuremath{\chi_{c0}}\xspace$, and NR components. We then add another scalar resonance described by the RBW parametrization.
We scan the likelihood by fixing the width and mass of this additional resonance at several consecutive values, for each of which
the fit to the data is repeated.
All isobar magnitudes and phases are floating in these fits.
From the scans we observe a significant improvement of the fit around a width and mass that are compatible with the values of the $\fV$ resonance~\cite{Amsler:2008zz}.
After adding the $\fV$ to the nominal model we repeat the same procedure for an additional tensor particle. We find that the $\fVII$ has a significant contribution. The results of the likelihood scans are shown in Fig.~\ref{fig:ModelScans} in terms of $-2 \Delta {\ln}{\cal L}=-2\rm\ln{\cal L}-(-2\ln{\cal L})_{min}$, where $(-2\ln{\cal L})_{\rm min}$ corresponds to the minimal value obtained in the particular scan. To conclude the search for possible resonant contributions we add all well established resonances~\cite{Amsler:2008zz} and check if the likelihood increases. We do not find any other significant resonant contribution but as we cannot exclude small contributions from the $f_{0}(1370)$, $f_{2}(1270)$, $f_{2}^{'}(1525)$, $a_{0}(1450)$, and $f_{0}(1500)$ resonances, we assign model uncertainties (see Sec.~\ref{sec:dp_systematics}) due to not taking these resonances into account.
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=8.0cm,keepaspectratio]{figures/ScanScalarOnBaseLine.eps}
\includegraphics[width=8.0cm,keepaspectratio]{figures/ScanTensorOnBasePlusf01710.eps}
\caption{
Two-dimensional scans of $-2 \Delta \ln{\cal L}$
(gray scale)
as a function of the mass and the width of an additional resonance. These scans were performed to look for an additional scalar resonance (left) and an additional tensor resonance (right). The baseline model of the scans for additional scalar resonances contains $\fI$, $\ensuremath{\chi_{c0}}\xspace$, and NR intermediate states. The baseline model of the scans for additional tensor resonances contains $\fI$, $\ensuremath{\chi_{c0}}\xspace$, NR, and $\fV$ intermediate states. The
ellipses indicate the world average parameters~\cite{Amsler:2008zz} for the $\fV$ and $\fVII$ resonances that are added to the model.}
\label{fig:ModelScans}
\end{center}
\end{figure*}
\subsection{Systematic uncertainties}
\label{sec:dp_systematics}
Systematic effects are divided into model and experimental uncertainties. Details on how they have been estimated are given below and the associated numerical values are summarized in Table~\ref{tab:systematics}.
\subsubsection{Model uncertainties}
We vary the mass, width, and any other parameter of all isobar fit components within their errors, as quoted in Table~\ref{tab:model}, and assign the observed differences in our observables as the first part of the model uncertainty (``Model'' in Table~\ref{tab:systematics}).
To estimate the contribution to $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ from resonances that are not included in our signal model but cannot be excluded statistically, namely the $f_{0}(1370)$, $f_{2}(1270)$, $f_{2}^{'}(1525)$, $a_{0}(1450)$, and $f_{0}(1500)$ resonances, we perform fits to pseudoexperiments that include these resonances. The masses and the widths are taken from~\cite{Amsler:2008zz}, except for the $f_{0}(1370)$ for which we take the values from~\cite{pappagallo}. We generate pseudoexperiments with the additional resonances, where the isobar magnitudes and phases have been determined in fits to data, and fit these datasets with the nominal model. We assign the induced shift in the observables as a second part of the model uncertainty.
\subsubsection{Experimental systematic uncertainties}
To validate the analysis procedure, we perform fits on a large number of pseudoexperiments generated with the measured yields of signal events and continuum background. The signal events are taken from fully reconstructed MC that has been generated with the fit result to data. We observe small biases in the isobar magnitudes and phases. We correct for these biases by shifting the values of the parameters and assign to this procedure a systematic uncertainty, which corresponds to half the correction combined in quadrature with its error. This uncertainty accounts also for correlations between the signal variables, wrongly reconstructed events, and effects due to the limited sample size (``Fit Bias'' in Table~\ref{tab:systematics}).
From MC we estimate that there are six \ensuremath{B}\xspace\ background events in our data sample. To determine the bias introduced by these events, we add \ensuremath{B}\xspace\ background events from MC to our data sample, and fit it with the nominal model. We then assign the observed differences in the observables as a systematic uncertainty (``\ensuremath{B}\xspace-bkg'' in Table~\ref{tab:systematics}).
We assign a systematic uncertainty for all fixed PDF parameters by varying them within their uncertainties according to the covariance matrix.
We vary the histogram PDFs, i.e., the SDP PDF for continuum and the NN PDF for signal (``Discr.~Vars'' in Table~\ref{tab:systematics}).
The \mbox{$m_{\rm ES}$}\xspace dependence of the SDP PDF for continuum was found to be negligible.
We account for differences between simulation and data observed in the control sample $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}$ (``MC-Data'' in Table~\ref{tab:systematics}).
These differences were estimated by propagating the differences, in the control sample, between background-subtracted data and signal MC, into the fit PDFs.
For the branching fraction measurement, we assign a systematic uncertainty due to the error on the calculation of $N_{B\bar B}$ (``$N_{B \bar B}$'' in Table~\ref{tab:systematics}) and to the $\ensuremath{K^0_S}$ reconstruction efficiency. We correct the $\ensuremath{K^0_S}$ reconstruction efficiency by the difference between the efficiency found in a dedicated \ensuremath{K^0_S}\ data sample and that found in simulation.
We assign the uncertainty on the correction as a systematic error (``\ensuremath{K^0_S}\ reco'' in Table~\ref{tab:systematics}).
\begin{table*}[htbp]
\caption{Summary of systematic uncertainties. The model uncertainty is dominated by the variation of the line shapes due to the contribution of the poorly measured $f_{2}(2010)$.}
\begin{tabular}{lccccccccc}
\hline \hline
Parameter &Fit bias&\ensuremath{B}\xspace-bkg & Discr.~Vars &MC-Data&$N_{B \bar B}$&$\ensuremath{K^0_S}$ reco& Sum & Model \\
${\cal B}(\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S})\,[10^{-6}]$ & $0.011$ & $0.030$ & $0.053$ & $0.015$ & $0.067$ & $0.111$ & $0.145$ & $0.120$ \\ \hline
FF \fI & $0.013$ & $0.056$ & $0.006$ & $0.001$ & - & - & $0.058$ & $0.190$ \\
FF \fV & $0.007$ & $0.001$ & $0.001$ & $0.001$ & - & - & $0.007$ & $0.016$ \\
FF $\fVII$ & $0.005$ & $0.001$ & $0.003$ & $0.001$ & - & - & $0.006$ & $0.084$ \\
FF NR & $0.024$ & $0.083$ & $0.023$ & $0.001$ & - & - & $0.090$ & $0.344$ \\
FF $\ensuremath{\chi_{c0}}\xspace$ & $0.002$ & $0.000$ & $0.001$ & $0.000$ & - & - & $0.002$ & $0.034$ \\
Ph [rad] \fI & $0.008$ & $0.018$ & $0.014$ & $0.000$ & - & - & $0.024$ & $0.177$ \\
Ph [rad] \fV & $0.011$ & $0.020$ & $0.001$ & $0.003$ & - & - & $0.023$ & $0.185$ \\
Ph [rad] \fVII & $0.044$ & $0.014$ & $0.004$ & $0.002$ & - & - & $0.046$ & $0.684$ \\
Ph [rad] $\ensuremath{\chi_{c0}}\xspace$ & $0.039$ & $0.011$ & $0.010$ & $0.007$ & - & - & $0.042$ & $0.498$ \\ \hline\hline
\end {tabular}
\label{tab:systematics}
\end{table*}
\section{THE \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ DETECTOR AND DATASET}
\label{sec:babar}
The data used in this analysis were collected with the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\
detector at the PEP-II\ asymmetric-energy $e^+e^-$ storage ring at
SLAC.
The sample consists of
an integrated luminosity of
$426.0\;\mathrm{fb}^{-1}$, corresponding to $(467.8\pm 5.1)\times10^{6}$
$B\kern 0.18em\overline{\kern -0.18em B}{}\xspace$ pairs collected at the \Y4S resonance (``on-resonance''),
and $44.5$~\ensuremath{\mbox{\,fb}^{-1}}\xspace collected about $40$~\ensuremath{\mathrm{\,Me\kern -0.1em V}}\xspace
below the~\Y4S (``off-resonance'').
A detailed description of the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ detector is presented in
Ref.~\cite{babar}. The tracking system used for track and vertex
reconstruction has two components: a silicon vertex tracker
and a drift chamber, both operating within a 1.5~T
magnetic field generated by a superconducting solenoidal magnet.
A detector of internally reflected Cherenkov light
associates Cherenkov photons with tracks for particle
identification.
The energies of photons and electrons
are determined from the measured light produced
in electromagnetic showers inside a CsI crystal electromagnetic calorimeter.
Muon
candidates are identified with the
use of the instrumented flux return of the solenoid.
\section{SUMMARY}
\label{sec:Summary}
We have performed the first amplitude analysis of $\ensuremath{B^0}\xspace \rightarrow \ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ events and measured the total inclusive branching fraction to be $\rm (6.19 \pm 0.48 \pm 0.15 \pm 0.12) \times 10^{-6}$, where the first uncertainty is statistical, the second is systematic, and the third represents the signal DP-model dependence. We have identified the dominant contributions to the DP to be from $\fI$, $\fV$, $\fVII$, and a nonresonant component, and measured the individual fit fractions and phases of each component. We do not observe any significant contribution from the so-called $f_X(1500)$ resonance seen in, for example, $\ensuremath{\Bu}\xspace\rightarrow \ensuremath{K^+}\xspace\ensuremath{K^-}\xspace\ensuremath{K^+}\xspace$~\cite{babarKKK}. The peak in the invariant mass between $1.5$ and $1.6\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ can be described by the interference between the $f_0(1710)$ resonance and the nonresonant component. We see some hints from the $f_{2}^{'}(1525)$ and $f_{0}(1500)$ resonances that could also contribute to this structure, but due to limited sample size we cannot make a significant statement. Future investigations of the $KK$ system could shed more light on the situation.
Furthermore we have performed an update of the phase-space-integrated time-dependent analysis of the same decay mode, using \pippim\ and \pizpiz\ decays, with the final \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ dataset. We measure the \ensuremath{C\!P}\xspace-violation parameters to be ${\cal S} = -0.94\,^{+0.24}_{-0.21} \pm 0.06 $ and ${\cal C} = -0.17 \pm 0.18 \pm 0.04 $, where the first quoted uncertainty is statistical and the second is systematic.
These measured values are consistent with and supersede those reported in Ref.~\cite{KsKsKsBabar}. They are compatible within two standard deviations with those measured in tree-dominated modes such as $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}$, as expected in the SM. For the first time, we report evidence of \ensuremath{C\!P}\xspace\ violation in $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ decays; \ensuremath{C\!P}\xspace\ conservation is excluded at $3.8$ standard deviations including systematic uncertainties.
\subsection{Event selection and backgrounds}
\label{sec:td_selection}
We reconstruct $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ candidates either from three $\ensuremath{K^0_S}\ensuremath{\rightarrow}\xspace\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$ candidates, or from two $\ensuremath{K^0_S}\ensuremath{\rightarrow}\xspace\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$ and one $\ensuremath{K^0_S}\ensuremath{\rightarrow}\xspace\ensuremath{\pi^0}\xspace\piz$, where the $\ensuremath{\pi^0}\xspace$ candidates are formed from pairs of photons. The vertex fit requirements are the same as in the amplitude analysis, and also the requirement that the charged pions of at least one of the $\ensuremath{K^0_S}$ have hits in the two inner layers of the vertex tracker.
The \ensuremath{K^0_S}\ candidates in the \pippim\ submode must have mass within $12\ensuremath{{\mathrm{\,Me\kern -0.1em V\!/}c^2}}\xspace$ of the nominal $\ensuremath{K^0}\xspace$ mass~\cite{Amsler:2008zz} and decay length with respect to the $B$ vertex between $0.2$ and $40\ensuremath{{\rm \,cm}}\xspace$.
In addition, combinatorial background is suppressed in both submodes by imposing that the angle between
the momentum vector of each $\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ candidate
and the vector connecting the beamspot and the $\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ vertex is smaller than 0.2~radians.
Each \ensuremath{K^0_S}\ decaying to charged pions in the \pizpiz\ submode
is required to have decay length between $0.15$ and $60$~cm and $\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$ invariant mass less than $11\ensuremath{\mathrm{\,Me\kern -0.1em V}}\xspace$ from the world average \ensuremath{K^0_S}\ mass~\cite{Amsler:2008zz}. The \ensuremath{K^0_S}\ decaying to neutral pions in the \pizpiz\ submode must have $\ensuremath{\pi^0}\xspace\piz$ invariant mass between $0.48$ and $0.52\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$. Additionally, the neutral pions are selected if they have $\gamma\gamma$ invariant mass between $0.100$ and $0.141\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ and if the photons have energies greater than $50\ensuremath{\mathrm{\,Me\kern -0.1em V}}\xspace$ in the laboratory frame and a lateral energy deposition profile in the electromagnetic calorimeter consistent with that expected for an electromagnetic shower (lateral moment~\cite{Drescher:1984rt} less than $0.55$).
The fact that we do not model any PDF using sideband data allows a loose requirement on $\mbox{$m_{\rm ES}$}\xspace$ and $\Delta E$ in the time-dependent analysis, namely, $5.22 < \mbox{$m_{\rm ES}$}\xspace <5.29\ensuremath{{\mathrm{\,Ge\kern -0.1em V\!/}c^2}}\xspace$ and $-0.18<\Delta E<0.12\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$.
In case of multiple candidates passing the selection, we proceed in the same way as in the amplitude analysis.
We use the same NN as in the amplitude analysis to suppress continuum background.
With the above selection criteria, we obtain signal reconstruction efficiencies of $6.7\%$ and $3.1\%$ for the \pippim\ and \pizpiz\ submodes, respectively. These efficiencies are determined from a DP-model MC sample generated using the results of the amplitude analysis.
We estimate from MC that $2.1\%$ of the selected signal events are misreconstructed for \pippim, while the figure is $2.4\%$ in \pizpiz,
and we do not treat these events differently from correctly reconstructed events.
Because of the looser requirements, there are more background events from \ensuremath{B}\xspace decays than in the amplitude analysis, in particular in the \pizpiz\ submode. These backgrounds are included in the fit model and are summarized in Table~\ref{tab:bbackground}.
As the analysis is phase-space integrated, we cannot model the $\ensuremath{\chi_{c0}}\xspace$ resonance separately, and its contribution to the \ensuremath{C\!P}\xspace\ asymmetries could cloud deviations in the charmless contributions.
We therefore apply a veto around the invariant mass of this charmonium state.
\begin{table*}[htbp]
\begin{center}
\caption{ \label{tab:bbackground} Summary of \ensuremath{B}\xspace background modes included in the fit model of the time-dependent analysis. The expected number of events takes into account the branching fractions (${\cal B}$) and efficiencies. In case there is no measurement, the branching fraction of an isospin-related channel is used. All the fixed yields are varied by $\pm 100\%$ for systematic uncertainties.}
\setlength{\tabcolsep}{0.0pc}
\begin{tabular*}{\textwidth}{@{\extracolsep{\fill}}lcccc}
\hline\hline
Submode & Background mode & Varied & ${\cal B}$ [$\times 10^{-6}$] & Number of events \\ \hline\\[-9pt]
\pippim\ & $\ensuremath{K^0_S}\KS\KL$ & no & $2.4$ & $ 0.71 $ \\
& $\ensuremath{K^0_S}\KS\ensuremath{K^{*0}}\xspace$ & no & $27.5$ & $ 9.55 $ \\
& $\ensuremath{K^0_S}\KS\ensuremath{K^+}\xspace$ & no & $11.5$ & $ 4.27 $ \\
& $\ensuremath{B^0}\xspace \rightarrow \{\text{neutral generic decays}\}$ & yes & not applicable & $21.7$ \\
& $\ensuremath{\Bu}\xspace \rightarrow \{\text{charged generic decays}\}$ & yes & not applicable & $15.5$ \\
\hline\\[-9pt]
\pizpiz\ & $\ensuremath{K^0_S}\KS\KL$ & no & $2.4 $ & $0.67 $ \\
& $\ensuremath{K^0_S}\KS\ensuremath{K^{*0}}\xspace$ & no & $27.5 $ & $5.3 $ \\
& $\ensuremath{K^0_S}\KL\ensuremath{K^{*0}}\xspace$ & no & $27.5 $ & $0.3 $ \\
& $\ensuremath{K^0_S}\KS\ensuremath{K^+}\xspace$ & no & $11.5 $ & $2.9 $ \\
& $\ensuremath{K^0_S}\KS\ensuremath{K^{*+}}\xspace$ & no & $27.5 $ & $7.2 $ \\
& $\ensuremath{B^0}\xspace \rightarrow \{\text{neutral generic decays}\}$ & yes & not applicable & $73.6$ \\
& $\ensuremath{\Bu}\xspace \rightarrow \{\text{charged generic decays}\}$ & yes & not applicable & $73.8$ \\
\hline\hline
\end{tabular*}
\end{center}
\end{table*}
\subsection{Results}
\label{sec:TD_fitresults}
The maximum-likelihood fit of $3261$ candidates in the $\pippim$ submode and $7209$ candidates in the $\pizpiz$ submode results in the event yields detailed in Table~\ref{tab:td_yields}.
\begin{table}[htb]
\begin{center}
\caption{ \label{tab:td_yields} Event yields for the different event species, resulting from the maximum-likelihood fit for the time-dependent analysis. ``$\ensuremath{\Bu}\xspace\ensuremath{\Bub}\xspace$ ($\ensuremath{B^0}\xspace\ensuremath{\Bbar^0}\xspace$) bkg'' represents background from charged (neutral) \ensuremath{B}\xspace decays. Quoted uncertainties are statistical only.}
\begin{tabular}{ lcc}
\hline \hline \noalign{\smallskip}
Species & $\ 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ $ & $\ 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)\ $\\
\noalign{\smallskip}\hline\noalign{\smallskip}
Signal & $\rm 201\,^{+16}_{-15}$ & $\rm 62\,^{+13}_{-12}$ \\[0.5mm]
Continuum & $\rm 3086\,^{+56}_{-54} $ & $\rm 7086\,^{+85}_{-83} $ \\[0.5mm]
$\ensuremath{\Bu}\xspace\ensuremath{\Bub}\xspace$ bkg & $\rm -54\,^{+29}_{-24}$ & $\rm 45\,^{+34}_{-30}$ \\[0.5mm]
$\ensuremath{B^0}\xspace\ensuremath{\Bbar^0}\xspace$ bkg & $\rm 9\,^{+31}_{-30}$ & $\rm 4\,^{+38}_{-29}$ \\[0.7mm] \hline \hline
\end{tabular}
\end{center}
\end{table}
The fit result for the time-dependent \ensuremath{C\!P}\xspace-violation parameters ${\cal S}$ and ${\cal C}$ is
\begin{eqnarray}
\nonumber {\cal S} &=& -0.94\,^{+0.24}_{-0.21}, \\
\nonumber {\cal C} &=& -0.17 \pm 0.18,
\end{eqnarray}
where the uncertainties are statistical only. The correlation between ${\cal S}$ and ${\cal C}$ is $-0.16$.
We use the fit result to create {\ensuremath{\hbox{$_s$}{\cal P}lots}} of the signal distributions of $\ensuremath{{\rm \Delta}t}\xspace$, the time-dependent asymmetry, and the discriminating variables. Figure~\ref{fig:Deltat_sPlots_TD} shows the $\ensuremath{{\rm \Delta}t}\xspace$ {\ensuremath{\hbox{$_s$}{\cal P}lots}} for the combined fit result and for the individual submodes.
Figure~\ref{fig:Discrimvars_sPlots_TD} shows the signal distributions and Fig.~\ref{fig:Discrimvars_Cont_sPlots_TD} the continuum background distributions of the discriminating variables.
The distributions shown in these three figures illustrate the good agreement between the data and the fit model.
We scan the statistical-only likelihood of the ${\cal S}$ parameter for both submodes and for the combined fit. The result, on the left-hand side of Fig.~\ref{fig:likelihoodScans_TD}, shows a sizable difference between the ${\cal S}$ values for the two submodes;
the level of consistency, conservatively estimated from the sum of the two individual likelihood scans, is approximately $2.6\sigma$ (a $p$-value of $1.0\%$ with $2$ degrees of freedom).
This value is obtained including only the dominant statistical uncertainty and neglecting the small correlation between the \ensuremath{C\!P}\xspace-violation parameters.
The results obtained when ${\cal S}$ and ${\cal C}$ are allowed to vary individually for each of the submodes are ${\cal S}=-1.42\,^{+0.27}_{-0.24}$, ${\cal C}=-0.14\,^{+0.17}_{-0.17}$ for $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ and ${\cal S}=0.40\,^{+0.56}_{-0.57}$, ${\cal C}=0.19\,^{+0.42}_{-0.43}$ for $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$. In both cases the quoted uncertainties are statistical only.
As there is some correlation between the ${\cal S}$ and ${\cal C}$ parameters, we perform a two-dimensional statistical likelihood scan of the combined likelihood, which is then convolved by the systematic uncertainties on ${\cal S}$ and ${\cal C}$ (systematic uncertainties are discussed in Sec.~\ref{sec:td_systematics}.)
The result is shown on the right-hand side of Fig.~\ref{fig:likelihoodScans_TD}.
We find that \ensuremath{C\!P}\xspace\ conservation is excluded at $3.8$ standard deviations, and thus, for the first time, we measure an evidence of \ensuremath{C\!P}\xspace\ violation in $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ decays.
The difference between our result and that from $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace c \bar c K^{(*)}$ is less than $2$ standard deviations.
The scan also shows that the result is close to the physical boundary, given by the constraint ${\cal S}^2+{\cal C}^2 \leq 1$.
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dt_TD_pippim_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dtAsym_TD_pippim_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dt_TD_pi0pi0_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dtAsym_TD_pi0pi0_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dt_TD_both_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_dtAsym_TD_both_June24.eps}
\end{center}
\caption{(color online). Signal {\ensuremath{\hbox{$_s$}{\cal P}lots}} (points with error bars) and PDFs (histograms) of $\ensuremath{{\rm \Delta}t}\xspace$ (left) and the derived asymmetry (right) for the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ submode (top), the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$ submode (middle), and the combined fit (bottom). In the $\ensuremath{{\rm \Delta}t}\xspace$ distributions on the left-hand side, points marked with $\times$ and solid lines correspond to decays where $\ensuremath{B}\xspace_{\rm tag}$ is a $\ensuremath{B^0}\xspace$ meson; points marked with $\circ$ and dashed lines correspond to decays where $\ensuremath{B}\xspace_{\rm tag}$ is a $\ensuremath{\Bbar^0}\xspace$ meson.
Points of the asymmetry {\ensuremath{\hbox{$_s$}{\cal P}lots}} that are outside the range of a figure are marked by arrows.}
\label{fig:Deltat_sPlots_TD}
\end{figure*}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_mES_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_mES_pi0pi0_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_DeltaE_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_DeltaE_pi0pi0_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_NN_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Signal_NN_pi0pi0_TD_June24.eps}
\end{center}
\caption{
(color online). Signal {\ensuremath{\hbox{$_s$}{\cal P}lots}} (points with error bars) and PDFs (histograms) of the discriminating variables: $\mbox{$m_{\rm ES}$}\xspace$ (top), $\Delta E$ (middle), and the NN output (bottom) for the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ submode (left) and for the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$ submode (right). Below each bin are shown the residuals, normalized in error units. The horizontal dotted and full lines mark the one- and two-standard deviation levels, respectively.}
\label{fig:Discrimvars_sPlots_TD}
\end{figure*}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_mES_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_mES_pi0pi0_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_DeltaE_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_DeltaE_pi0pi0_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_NN_pippim_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_NN_pi0pi0_TD_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_dt_TD_pippim_June24.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/sPlot_Cont_dt_TD_pi0pi0_June24.eps}
\end{center}
\caption{(color online). Continuum {\ensuremath{\hbox{$_s$}{\cal P}lots}} (points with error bars) and PDFs (histograms) of $\mbox{$m_{\rm ES}$}\xspace$, $\Delta E$, the NN output, and $\ensuremath{{\rm \Delta}t}\xspace$ (top to bottom). Plots on the left-hand side correspond to the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ submode, and on the right-hand side to the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$ submode. In the $\ensuremath{{\rm \Delta}t}\xspace$ distributions, points marked with $\times$ and solid lines correspond to decays where $\ensuremath{B}\xspace_{\rm tag}$ is a $\ensuremath{B^0}\xspace$ meson; points marked with $\circ$ and dashed lines correspond to decays where $\ensuremath{B}\xspace_{\rm tag}$ is a $\ensuremath{\Bbar^0}\xspace$ meson. Below each bin are shown the residuals, normalized in error units. The horizontal dotted and full lines mark the one- and two-standard deviation levels, respectively.}
\label{fig:Discrimvars_Cont_sPlots_TD}
\end{figure*}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=7.0cm,keepaspectratio]{figures/Sscan_1dim.eps}
\includegraphics[width=7.0cm,keepaspectratio]{figures/SCscan_2dim.eps}
\end{center}
\caption{(color online). One-dimensional statistical scan of $-2 \Delta \ln{\cal L}$ as a function of ${\cal S}$ (left) and the two-dimensional scan, including systematic uncertainty, as a function of ${\cal S}$ and ${\cal C}$ (right). In the left-hand plot, red points marked with $\times$ correspond to the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 3\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ submode, blue points marked with $\circ$ to the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace 2\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$ submode, and black points marked with $\ast$ to the combined fit. In the right-hand plot,
the gray scale is given in units of $\sqrt{-2 \Delta \ln{\cal L}}$.
The result of the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ analyses of $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace c \bar c K^{(*)}$ decays~\cite{LastSin2betaBabar} is indicated as a white ellipse and the physical boundary (${\cal S}^2+{\cal C}^2 \leq 1$) is marked as a gray line. The scan appears to be trimmed on the lower left since the PDF becomes negative outside the physical region (i.e., the white region does not indicate that the scan flattens out at 5$\sigma$).}
\label{fig:likelihoodScans_TD}
\end{figure*}
\subsection{The maximum-likelihood fit}
\label{sec:td_likelihood}
We perform an unbinned extended maximum-likelihood fit to extract the $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ event yields along with
the ${\cal S}$ and ${\cal C}$ parameters of the time-dependent analysis.
The fit uses as variables $\mbox{$m_{\rm ES}$}\xspace$, $\Delta E$, the NN output, $\ensuremath{{\rm \Delta}t}\xspace$, and $\sigma_{\ensuremath{{\rm \Delta}t}\xspace}$. The selected on-resonance data sample is assumed to consist of signal, continuum background, and backgrounds from $B$ decays. Wrongly reconstructed signal events are not considered separately. The likelihood function ${\cal L}_i$ for event $i$ is the sum
\begin{equation}
{\cal L}_i = \sum_j{N_j {\cal P}^i_j(\mbox{$m_{\rm ES}$}\xspace,\Delta E,\Delta t,\sigma_{\ensuremath{{\rm \Delta}t}\xspace},{\rm NN};q_{\rm tag},c,p)},
\end{equation}
where $j$ stands for the species (signal, continuum background, one for each $\ensuremath{B}\xspace$ background category) and $N_j$ is the corresponding yield; $q_{\rm tag}$, $c$, and $p$ are the tag flavor, the tagging category, and the physics category, respectively.
To determine $q_{\rm tag}$ and $c$ we use the \ensuremath{B}\xspace\ flavor-tagging
algorithm of Ref.~\cite{BabarS2b}. This algorithm combines several
different signatures, such as charges, momenta, and decay
angles of charged particles in the event to achieve optimal
separation between the two \ensuremath{B}\xspace\ flavors. This produces six
mutually exclusive tagging categories.
We also retain untagged events in a seventh category;
although these events do not contribute to the measurement
of the time-dependent CP asymmetry they do provide
additional sensitivity for the measurement of direct \ensuremath{C\!P}\xspace\
violation~\cite{Gardner:2003su}.
The two physics categories correspond to \pippim\ and \pizpiz\ decays.
The PDF for species $j$ evaluated for event $i$ is given by the product of individual PDFs:
\begin{align}
&{\cal P}^i_j(\mbox{$m_{\rm ES}$}\xspace,\Delta E,\Delta t,\sigma_{\ensuremath{{\rm \Delta}t}\xspace},{\rm NN};q_{\rm tag},c,p) = \\
\nonumber & {\cal P}^i_j(\mbox{$m_{\rm ES}$}\xspace;p)\:{\cal P}^i_j(\Delta E;p)\:{\cal P}^i_j({\rm NN};c,p)\:{\cal P}^i_j(\Delta t,\sigma_{\ensuremath{{\rm \Delta}t}\xspace};q_{\rm tag},c,p).
\end{align}
To take into account the different reconstruction of the two submodes,
we use separate PDFs for the two physics categories. Separate NN and
$\ensuremath{{\rm \Delta}t}\xspace$ PDFs are included for each tagging category within each physics
category. The separate $\ensuremath{{\rm \Delta}t}\xspace$ PDFs for the two physics categories allow
us to fit the
${\cal S}$ and ${\cal C}$ parameters either separately for the two submodes, or together.
The total likelihood is given by
\begin{equation}
\label{eq:td_Likelihood}
{\cal L}=\exp(-\sum_j N_j)\prod_i {\cal L}_i.
\end{equation}
\subsubsection{$\Delta t$ PDFs}
\label{sec:td_likeDalitz}
The signal PDF for $\Delta t$ is given in Eq.~\eqref{eq:dtmeas}.
Parameters that depend solely on the tag side of the events (namely, ${\langle D \rangle}_c$ and $\Delta D_c$) are taken from the analysis of $\ensuremath{B}\xspace \ensuremath{\rightarrow}\xspace c \bar c K^{(*)}$ decays~\cite{LastSin2betaBabar}. On the other hand, parameters that depend on the signal-side reconstruction, due to the absence of direct tracks from the \ensuremath{B}\xspace decay, cannot be taken from modes that include such direct tracks. This is the case for the parameters that describe the resolution function, which are found in a fit to simulated events. A systematic uncertainty for data-MC differences is assigned using the control sample $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}$, as explained in Sec.~\ref{sec:td_systematics}.
For continuum events we use a zero-lifetime component. This parametrization is convolved with the
same resolution function as for signal, with different parameters that are varied in the fit to data.
The parameters of this PDF are not separated in the tagging categories.
The small contribution from $\ensuremath{e^+e^-}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{c\overline c}\xspace$ events is well described by the tails of the resolution function.
For \ensuremath{B}\xspace\ background events we use the signal PDF, with resolution parameters
from the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ $\ensuremath{B}\xspace \ensuremath{\rightarrow}\xspace c\bar c K^{(*)}$ analysis~\cite{LastSin2betaBabar}.
The parameters ${\cal S}$ and ${\cal C}$ are set to zero and varied
to assign a systematic uncertainty.
\subsubsection{Description of the other variables}
\label{sec:td_likeDiscrim}
The \mbox{$m_{\rm ES}$}\xspace\ and \mbox{$\Delta E$}\xspace\ distributions of signal events are parametrized by an asymmetric Gaussian with power-law tails, as given in Eq.~\eqref{eq:cruijff}, and, for \mbox{$m_{\rm ES}$}\xspace, a small additional component, parameterized by an ARGUS shape function~\cite{argus}, to correctly describe misreconstructed events. The means in these two PDFs for \pippim\ events are allowed to vary in the fit to data, and the other parameters are taken from MC simulation.
For the NN distributions of signal we use histogram PDFs taken from MC simulation for each physics and tagging category.
The \mbox{$m_{\rm ES}$}\xspace, $\mbox{$\Delta E$}\xspace$, and NN PDFs for continuum events are parametrized by an ARGUS shape function, a straight line and the sum of power functions from Eq.~\eqref{eq:spexp}, respectively. All continuum parameters, except for
$c_2$ and $c_3$
of the NN PDF, are allowed to vary in the fit. All the fixed parameters are varied, within the uncertainties found in a fit to sideband data, to estimate systematic errors.
All the $B$ background PDFs are described by fixed histograms taken from MC simulation.
\subsection{Systematic uncertainties}
\label{sec:td_systematics}
The systematic uncertainties are summarized in Table~\ref{tab:sys_TD}. The ``${\rm MC_{stat}}$'' uncertainty accounts for the limited size of the simulated data samples used to create the PDFs. The ``$\ensuremath{B}\xspace_{\rm reco}$'' uncertainty propagates the experimental uncertainty in the measurement of tag-side-related quantities taken from~\cite{LastSin2betaBabar} to our measurement. The ``\ensuremath{B}\xspace-bkg'' contribution results from the uncertainty in the \ensuremath{C\!P}\xspace content and the branching fractions of fixed yields in the model of background from \ensuremath{B}\xspace decays. The dominant ``MC-Data: \ensuremath{{\rm \Delta}t}\xspace'' systematic uncertainty is due to possible differences between data and simulation
concerning the procedure used to obtain the signal \ensuremath{B}\xspace decay vertex from tracks originating from $\ensuremath{K^0_S}$ decays.
We quantify this uncertainty using the control sample $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}$ by comparing the difference between $\Delta t$ values obtained with and without the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace$ in data and simulation. We then propagate the observed differences and their uncertainties to the resolution function. We use this new resolution function to refit the data and obtain an estimate of the effect on ${\cal S}$ and ${\cal C}$. We also use the samples $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}(\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace)$ and $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}(\ensuremath{\pi^0}\xspace\piz)$ to estimate simulation-data differences for the other variables in the submodes \pippim\ and \pizpiz, respectively. This contribution is referred to as ``MC-Data: Discr.~Vars'' in Table~\ref{tab:sys_TD}. The ``Fit Bias'' uncertainty is evaluated using fits to fully reconstructed MC samples. It accounts for effects from wrongly reconstructed events and correlations between fit variables. The ``Vetoes'' uncertainty is related to the veto on the invariant mass. It is evaluated using events that pass the veto in pseudo-experiments studies. Finally, the ``Misc'' uncertainty includes contributions from doubly-Cabibbo-suppressed decays, silicon vertex tracker alignment, and the uncertainties in the boost of the $\Y4S$. These contributions are taken from the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ $\ensuremath{B}\xspace \ensuremath{\rightarrow}\xspace c\bar c K^{(*)}$ analysis~\cite{LastSin2betaBabar}.
\begin{table*}[htbp]
\begin{center}
\caption{Summary of systematic uncertainties on the ${\cal S}$ and ${\cal C}$ parameters.
\label{tab:sys_TD}}
\begin{tabular}{lcc}
\hline \hline
Source & ${\cal S}$ & ${\cal C}$ \\ \hline
$\rm MC_{\rm stat}$ & $0.002$ & $0.001$ \\
$\ensuremath{B}\xspace_{\rm reco}$ & $0.004$ & $0.003$ \\
$\ensuremath{B}\xspace$-bkg & $0.032$ & $0.012$ \\
MC-Data: \ensuremath{{\rm \Delta}t}\xspace & $0.045$ & $0.027$ \\
MC-Data: Discr.~Vars & $0.021$ & $0.004$ \\
Fit Bias & $0.022$ & $0.018$ \\
Vetoes & $0.006$ & $0.004$ \\
Misc & $0.004$ & $0.015$ \\ \hline
Sum & $0.064$ & $0.038$ \\ \hline \hline
\end{tabular}
\end{center}
\end{table*}
\section{Time-dependent analysis}
\label{sec:TD_analysis}
In Sec.~\ref{sec:dt} we describe the proper-time distribution used to extract the time-dependent \ensuremath{C\!P}\xspace\ asymmetries.
In Sec.~\ref{sec:td_selection} we explain the selection
requirements used to obtain the signal candidates and
suppress backgrounds. In Sec.~\ref{sec:td_likelihood} we describe the fit method
and the approach used to account for experimental effects.
In Sec.~\ref{sec:TD_fitresults} we present the results of the fit and finally, in Sec.~\ref{sec:td_systematics} we discuss systematic uncertainties in the results.
\subsection{Proper-time distribution}
\label{sec:dt}
The time-dependent \ensuremath{C\!P}\xspace\ asymmetries are functions of the proper-time difference $\ensuremath{{\rm \Delta}t}\xspace =t_{\ensuremath{C\!P}\xspace} - t_{\rm tag}$ between a fully reconstructed \ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}\ decay ($B_{\ensuremath{C\!P}\xspace}$) and the other $B$ meson decay in the event ($B_{\rm tag}$), which is partially reconstructed. The observed decay rate is the physical decay rate modified to include tagging imperfections, namely ${\langle D \rangle}_c$ and $\Delta D_c$; the former
is the rate of correctly assigning the flavor of the \ensuremath{B}\xspace\ meson, averaged over \ensuremath{B^0}\xspace\ and \ensuremath{\Bbar^0}\xspace,
and the latter is the difference between $D_c$ for \ensuremath{B^0}\xspace\ and \ensuremath{\Bbar^0}\xspace.
The index $c$ denotes different quality categories of the tag-flavor assignment.
Furthermore the decay rate is convolved with the per-event $\ensuremath{{\rm \Delta}t}\xspace$ resolution ${\cal R}_{\rm sig}(\ensuremath{{\rm \Delta}t}\xspace,\sigma_{\ensuremath{{\rm \Delta}t}\xspace})$, which is described by the sum of three Gaussians and depends on $\ensuremath{{\rm \Delta}t}\xspace$ and its error $\rm \sigma_{\ensuremath{{\rm \Delta}t}\xspace}$.
For an event $i$ with tag flavor $q_{\rm tag}$, one has
\begin{eqnarray}
\label{eq:dtmeas}
\lefteqn{{\cal P}^i_{\rm sig}(\Delta t,\sigma_{\ensuremath{{\rm \Delta}t}\xspace};q_{\rm tag},c) =}\\
\nonumber & & \frac{e^{-\left|\ensuremath{{\rm \Delta}t}\xspace\right|/\tau_{\ensuremath{B^0}\xspace}}}{4\tau_{\ensuremath{B^0}\xspace}} \biggl\{ 1 + q_{\rm tag}\frac{\Delta D_{c}}{2} \\
\nonumber & & + q_{\rm tag}\langle D\rangle_{c} \Bigl[{\cal S}\sin(\Delta m_d\ensuremath{{\rm \Delta}t}\xspace)-\;{\cal C}\cos(\Delta m_{d}\ensuremath{{\rm \Delta}t}\xspace)\Bigr] \biggr\} \\
\nonumber & & \;\otimes\; {\cal R}_{\rm sig}(\ensuremath{{\rm \Delta}t}\xspace,\sigma_{\ensuremath{{\rm \Delta}t}\xspace}), \nonumber
\end{eqnarray}
where $q_{\rm tag}$ is defined to be $+1$ ($-1$) for $B_{\rm tag}=\ensuremath{B^0}\xspace$ ($B_{\rm tag}=\ensuremath{\Bbar^0}\xspace$), $\tau_{\ensuremath{B^0}\xspace}$ is the mean \ensuremath{B^0}\xspace\ lifetime, and $\Delta m_{d}$ is the mixing frequency~\cite{Amsler:2008zz}. The widths of the $\ensuremath{B^0}\xspace$ and the $\ensuremath{\Bbar^0}\xspace$ are assumed to be the same.
\section{INTRODUCTION}
\label{sec:introduction}
Over the past ten years, the $B$ factories have shown that the
Cabibbo-Kobayashi-Maskawa paradigm in the standard model (SM),
with a single weak phase in the quark mixing matrix, accounts for the
observed \ensuremath{C\!P}\xspace-symmetry violation in the quark sector.
However, there may be other \ensuremath{C\!P}\xspace-violating sources beyond the SM.
Charmless hadronic $B$ decays, like $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$, are of great interest because they are dominated by loop diagrams and are thus sensitive to new physics effects at large energy scales~\cite{Grossman:1996ke}.
In the SM, the mixing-induced \ensuremath{C\!P}\xspace-violation parameters in this decay are expected to be the same, up to $\sim 1\%$~\cite{Cheng:2005ug}, as in the tree-diagram-dominated modes such as $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace\ensuremath{K^0_S}$.
Both \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}~\cite{KsKsKsBabar} and Belle~\cite{belleCP} have previously performed time-dependent \ensuremath{C\!P}\xspace-violation measurements of the inclusive mode $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$, which is permissible because the final state is \ensuremath{C\!P}\xspace\ definite~\cite{Gershon:2004tk}.
The structure of the Dalitz plot (DP), however, is of interest; although the
time-dependent \ensuremath{C\!P}\xspace-violation parameters ${\cal S}$ and ${\cal C}$ [see Eq.~\eqref{eq:dtmeas}] can be measured inclusively without taking into account the phase space, different resonant contributions may have different values of these parameters in the presence of new physics. The statistical precision is not sufficient to perform a time-dependent amplitude analysis, but as we show below, it is possible to extract branching fractions from resonant contributions to the decay using a time-integrated amplitude analysis.
Additionally, the amplitude analysis could shed light on the controversial \fXI\ resonance: recent measurements of $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace \ensuremath{K^+}\xspace\ensuremath{K^-}\xspace\ensuremath{K^0_S}$ and $\ensuremath{B^\pm}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{K^+}\xspace\ensuremath{K^-}\xspace\ensuremath{K^\pm}\xspace$ from \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ \cite{babarKKK, Aubert:2007sd, babarKKKs} and Belle~\cite{belleKKK, belleKKKs} have shown evidence of a wide structure in the $m_{\ensuremath{K^+}\xspace\ensuremath{K^-}\xspace}$ spectrum around $1.5\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$.
In these measurements, it was assumed that this structure is a single scalar resonance; however a vector hypothesis could not be ruled out. The \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ measurement of $\ensuremath{\Bu}\xspace \rightarrow \ensuremath{K^+}\xspace\ensuremath{K^-}\xspace \ensuremath{\pi^+}\xspace$ \cite{babarKKpi} appears to show an enhancement around $1.5\ensuremath{\mathrm{\,Ge\kern -0.1em V}}\xspace$, while the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ analysis of $\ensuremath{B^\pm}\xspace \rightarrow \ensuremath{K^0_S} \ensuremath{K^0_S} \ensuremath{\pi^\pm}\xspace$~\cite{KsKspi} finds no evidence of a possible \fXI, suggesting that the structure is either a vector meson or something exotic. An amplitude analysis of $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ will provide further insight into the nature of this structure, as only intermediate states of even spin are permitted due to Bose-Einstein statistics; an observation of the \fXI\ decaying to \ensuremath{K^0_S}\KS\ would require an even-spin state.
Finally, the amplitude analyses of $\ensuremath{B}\xspace \ensuremath{\rightarrow}\xspace K\pi\pi$ and $\ensuremath{B}\xspace \ensuremath{\rightarrow}\xspace KKK$ modes may be used to extract the Cabibbo-Kobayashi-Maskawa angle $\gamma$~\cite{London_Lorier:2011}.
This paper presents the first amplitude analysis and the final \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ update of the time-dependent \ensuremath{C\!P}\xspace-asymmetry measurement of $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ using the full \Y4S\ dataset.
The amplitude analysis is time-integrated \ensuremath{C\!P}\xspace-averaged (i.e., it does not use flavor-tagging information to distinguish between \ensuremath{B^0}\xspace\ and \ensuremath{\Bbar^0}\xspace mesons). It takes advantage of the interference pattern in the DP to measure relative magnitudes and phases for the different resonant modes using $\ensuremath{B^0}\xspace\ensuremath{\rightarrow}\xspace\ensuremath{K^0_S}\KS\ensuremath{K^0_S}$ decays with $\ensuremath{K^0_S}\ensuremath{\rightarrow}\xspace\ensuremath{\pi^+}\xspace\ensuremath{\pi^-}\xspace$, denoted by \pippim.
The magnitudes and phases are then translated into individual branching fractions for the resonant modes.
The time-dependent analysis extracts the ${\cal S}$ and ${\cal C}$ parameters by modeling the proper-time distribution. This part of the analysis uses both \pippim\ events and events where one of the \ensuremath{K^0_S}\ mesons decays to $\ensuremath{\pi^0}\xspace\piz$, denoted by \pizpiz.
In Sec.~\ref{sec:babar} we briefly describe the \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}\ detector
and the dataset. The amplitude analysis is described in Sec.~\ref{sec:DP_analysis} and the time-dependent analysis in Sec.~\ref{sec:TD_analysis}. Finally we summarize the results in Sec.~\ref{sec:Summary}.
\section*{ACKNOWLEDGMENTS}
\label{sec:acknowledgments}
\input{acknowledgements}
|
\section{Introduction}
The balanced homodyne detector (HD) is a useful tool in quantum optics and quantum information processing with continuous variables \cite{leo:97,lvov:09} since it can be used to measure field quadratures of an electromagnetic mode. These measurements provide information for complete reconstruction of quantum states in the optical domain (optical homodyne tomography).
With developing tools of continuous-wave quantum-optical state engineering \cite{bim:10} as well as state and process tomography \cite{lob:08}, the performance requirements for homodyne detectors continue to increase. The design of HDs for time-domain quantum tomography \cite{han:01, vogel:89, zav:02} is based on four main performance criteria: a) high bandwidth and a flat amplification profile within that bandwidth; b) high ratio of the measured quantum noise over the electronic noise; c) high common mode rejection ratio (CMRR); d) quantum efficiency of the photodiodes.
The \emph{high bandwidth} requirement comes from the fact that an HD must be able to measure field quadratures with sufficient time resolution. In the case of pulsed lasers, this corresponds to the inverse of the repetition rate of the pulses; in the case of a continuous signal, the required resolution is determined by the duration of the optical mode in which the signal states are produced \cite{sas:06, nie:09, yue:83}. This is technically challenging because most amplifiers have a limited gain-bandwidth product. Increasing the bandwidth implies reducing the gain, which, in turn, increases the effect of the electronic noise. Also the high frequency circuit layout poses a challenge to designers.
Within its bandwidth range, the HD must feature a \emph{flat amplification profile}. If this is not the case, the response of the HD to each individual pulse will exhibit ringing, which degrade the detector's time resolution and distort the measurement. This requirement also presents a major design challenge.
Any non-desirable ambient noises, dark current noises from the photodiodes and the intrinsic noise of the amplifiers fall under the umbrella of \emph{electronic noise}. The effect of this noise is to add a random quantity $Q_e$ to the measurement of the field quadrature $Q_{\rm meas}$. This effect is equivalent to an additional optical loss channel with transmission \cite{appel:07}
\begin{equation}\label{etae}
\eta_e=1-\langle\hat Q_e^2\rangle/\langle \hat Q_{\rm meas}^2\rangle.
\end{equation}
As we show below, the value of $\eta_e$ depends not only on the characteristics of the detector, but also on the conditions of the measurement in which the detector is used.
A HD must have a \emph{high subtraction capability} between the two photocurrents produced by the photodiodes. This can be expressed as a generalized common mode rejection ratio (CMRR) of the balanced detection \cite{vogel:89,chi:11}. The CMRR measures the ability of the device to reject the classical noise of the local oscillator \cite{bac:04,sas:06}.
This is particularly important in the pulsed case because a low CMRR (which implies a poor subtraction) will result in contamination of the signal with the repetition rate of the pulse and harmonics. Additionally, this lack of subtraction capability will make the HD more susceptible to saturation by the amplified signal from the photodiodes. High CMRR is difficult to achieve because the response functions of the photodiodes are not exactly the same. Therefore a pair of photodiodes with response functions as similar as possible must be chosen.
Experimentally, these performance benchmarks can be measured using an electronic spectrum analyzer. The spectrum of the homodyne output photocurrent gives information about the detector's bandwidth and amplification profile. Observing the output current in the absence of the local oscillator provides information about the magnitude and spectrum of the electronic noise. The lower bound on CMRR is determined by comparing the HD spectra when both photodiodes are illuminated and when only one is illuminated while the other is blocked.
In the present work, we quantitatively relate the measured electronic spectra to added noise in quadrature measurements. We show that the limited bandwidth and electronic noise can be translated into equivalent optical losses such as in \eeqref{etae}. We show how to estimate and reduce these losses for a specific time-domain experiment. In fact, in many cases (particularly, in the continuous-wave regime) electronic spectral measurements on the HD photocurrent in the presence and absence of the local oscillator are sufficient to precisely calculate the equivalent loss associated with the electronics.
The theoretical discussion in this paper is limited to the effects of the bandwidth and the electronic noise in two practically relevant regimes. The effect of the non-unitary quantum efficiency on quantum state reconstruction is well known \cite{leo:97}. A discussion of CMRR has been presented in detail in Ref.~\cite{chi:11}.
We then demonstrate an easy to follow method for the design and construction of a wideband homodyne detector using commercial available components such as low-noise high-speed operational amplifiers and high-bandwidth photodiodes. Aside from high performance benchmarks, a special feature of our detector is its versatility: it is designed and tested to operate in both the continuous-wave or pulsed regimes. Therefore the unit presented here may be useful for a wide range of quantum optics experiments.
\section{Theoretical analysis}
Balanced homodyne detection consists of overlapping the signal mode carrying the quantum state in question and a strong reference field in a matching mode (the local oscillator, or LO) on a symmetric beam splitter. The two output signals of this beam splitter are directed to the two photodiodes of the HD, where these fields are detected and subtracted. Neglecting experimental imperfections, the subtraction photocurrent is then
\begin{equation}\label{idealHD}
\hat i(t)=A\alpha(t)\hat q_\theta(t),
\end{equation}
where $\hat{q}_\theta(t)$ is the instantaneous field quadrature value in the signal mode, $\alpha(t)$ and $\theta$ are the local oscillator amplitude and phase, respectively, and $A$ is a proportionality coefficient related to the HD amplifier gain. It is assumed that the local oscillator phase is constant.
The instantaneous quadrature observable can be written as
\begin{equation}\label{qtheta}
\hat{q}_\theta(t)=\hat a(t)e^{i\theta}+\hat a^\dag(t)e^{-i\theta},
\end{equation}
where $\hat a(t)$ is the time dependent photon annihilation operator\cite{fed:05}.
In a practical HD, the relationship between the quadrature measurement and the output current is more complex. It can be approximated by
\begin{equation}\label{cpracHD}
\hat i(t)=\hat i_e(t)+A\intinf \alpha(t')\hat q(t')r(t-t')\de t',
\end{equation}
where $i_e(t)$ is the detector's electronic noise and $r(\cdot)$ is its response function. An ideal detector would have $i_e(t)=0$ and $r(\tau)=\delta(\tau)$. In practice these conditions are not met.
As evident from \eeqref{cpracHD}, the impact of the electronic noise is minimized by raising the power of the local oscillator and the amplifier gain. However, practical possibilities of increasing the gain without proportionally increasing the electronic noise are limited. The local oscillator power must also be restricted to avoid saturation of the photodiodes and eliminating the classical noise \cite{bac:04}. Therefore in the analysis below we assume $A$ and $\alpha$ to equal their optimal values for the given experimental setting.
\subsection{Continuous regime}
In the continuous regime, the amplitude of the local oscillator is a constant: $\alpha(t)\equiv\alpha$. We are interested in measuring the quadrature of the signal field associated with a particular (normalized) temporal mode function $\phi(t)$, which we assume real:
\begin{equation}\label{QCV}
\hat Q=\intinf \hat q(t)\phi(t)\de t
\end{equation}
To that end, we integrate the homodyne photocurrent with a certain weight function $\psi(t)$, obtaining a \emph{measured} quadrature value,
\begin{eqnarray}\label{QmeasCV}
\hat Q_{\rm meas}&=&\intinf \hat i(t)\psi(t)\de t \\ \nn
&=&A\alpha\intinf\intinf \hat q(t')\psi(t)r(t-t')\de t \de t'+\hat Q_e,
\end{eqnarray}
in which the last term,
\begin{equation}\label{}
\hat Q_e=\intinf \hat i_e(t)\psi(t)\de t,
\end{equation}
corresponds to the electronic noise contribution, which we will discuss later. First, we discuss the effect of finite detector response function (bandwidth) on the quadrature measurement.
Equation \eqref{QmeasCV} can be rewritten as
\begin{equation}\label{QmeasCV1}
\hat Q_{\rm meas}=A\alpha\intinf \hat q(t')\psi'(t')\de t'+\hat Q_e,
\end{equation}
where
\begin{equation}\label{}
\psi'(t')=\intinf \psi(t)r(t-t')\de t.
\end{equation}
By comparing Eqs.~\eqref{QCV} and \eqref{QmeasCV1} we find that, by choosing $\psi(t)$ such that $\psi'(t)=\phi(t)$, we have $\hat Q_{\rm meas}=A\alpha\hat Q+\hat Q_e$, i.e., the distortions associated with the detector's finite bandwidth are completely eliminated. This may however be difficult in practice, because the required weight function is a \emph{deconvolution} of the temporal mode of interest and the detector's response. Lack of precise knowledge of either of the above may lead to significant errors in deconvolving.
If $\psi'(t)\ne\phi(t)$, the detection efficiency is degraded by the mode matching factor \cite{aic:02}
\begin{equation}\label{MMCV}
\eta_b=\left.\left|\intinf\psi'(t)\phi(t)\de t\right|^2\middle/\intinf|\psi'(t)|^2\de t\right.,
\end{equation}
where the denominator normalizes $\psi'(t)$. A practically important particular case is when the temporal mode of the signal is known and the finite bandwidth of the detector is neglected, so $\psi(t)$ is set to equal $\phi(t)$. In Fig.~\ref{MMCVfig}, the efficiency \eqref{MMCV} obtained in this setting is plotted for Gaussian $\phi(t)$ and $r(t)$ as a function of the detector bandwidth, which, as we show below, is obtained from the Fourier transform of the response function.
As we see, the detector bandwidth has no significant degrading effect on the measurement ($\eta_b>0.99$) as long as it is comparable to or larger than the inverse temporal width of the signal temporal mode.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.8\columnwidth]{MMCV.eps}
\caption{Effective efficiency \eqref{MMCV} of the HD associated with its finite bandwidth. The 3-dB bandwidth is plotted along the horizontal axis in units of the inverse full-width-at-half-maximum of the signal temporal mode $\phi(t)$. Both functions are assumed Gaussian. \label{MMCVfig}}
\end{center}
\end{figure}
Let us now calculate the contribution of the electronic
noise to the measured quadrature, so the equivalent efficiency \eqref{etae} can be estimated. We start with the electronic spectra of the HD output photocurrent in the absence and in the presence of the local oscillator, with the signal in the vacuum state. According to the Wiener-Khintchine theorem, these spectra are given, respectively, by
\begin{equation}\label{}
S_e(\nu)=\intinf \langle \hat i_e(t)\hat i_e(t+\tau)\rangle e^{2\pi i\nu\tau}\de\tau
\end{equation}
and
\begin{equation}\label{Snu}
S(\nu)=\intinf \langle \hat i(t)\hat i(t+\tau)\rangle e^{2\pi i\nu\tau}\de\tau,
\end{equation}
where $\nu$ is the frequency and the averaging is performed over both time $t$ and the quantum ensemble of the vacuum signal state.
The autocorrelation function in the latter equation can be further simplified as
\begin{eqnarray}
\langle \hat i(t)\hat i(t+\tau)\rangle&& \\\nn &&\hspace{-1.8cm}
= A^2\alpha^2 \left\langle\intinf\intinf \hat q(t')r(t-t')\hat q(t'')r(t+\tau-t'')\de t'\de t''\right\rangle \\\nn
&&\hspace{-1cm}+\langle \hat i_e(t)\hat i_e(t+\tau)\rangle\\\nn
&&\hspace{-1.8cm}=A^2\alpha^2\intinf\langle r(t-t')r(t+\tau-t')\rangle\de t'+\langle \hat i_e(t)\hat i_e(t+\tau)\rangle\\\nn &&\hspace{-1.8cm}=A^2\alpha^2\intinf r(t)r(t+\tau)\de t+\langle \hat i_e(t)\hat i_e(t+\tau)\rangle
\end{eqnarray}
because, in the vacuum state, $\langle \hat q(t')\hat q(t'')\rangle=\delta(t'-t'')$. From the above, we find
\begin{equation}\label{SnuCV}
S(\nu)=A^2\alpha^2|\tilde r(\nu)|^2+S_e(\nu),
\end{equation}
with
\begin{equation}\label{}
\tilde r(\nu)=\intinf r(t) e^{2\pi i \nu\tau}\de t
\end{equation}
being the Fourier image of $r(t)$. In other words, neglecting the electronic noise, the spectrum of the HD output current in the continuous regime is simply the squared amplitude of the Fourier transform of the detector's response function. Note, however, that the response function cannot be obtained from this spectrum because inverse Fourier transformation also requires data on the phase of $\tilde r(\nu)$. The response function can be measured directly in the time domain with a pulsed LO as discussed below.
Let us now discuss the contribution of electronic noise to quadrature measurements. From \eeqref{QmeasCV1}, and because in the vacuum state $\langle \hat q(t)\hat q(t')\rangle=\delta(t-t')$, we can write
\begin{eqnarray}\label{QmeasCVvar}
\langle \hat Q_{\rm meas}^2\rangle &=& A^2\alpha^2\intinf |\psi'(t)|^2\de t + \langle \hat Q_e^2\rangle\\ \nn
&=& A^2\alpha^2\intinf |\tilde\psi'(\nu)|^2\de \nu+ \langle \hat Q_e^2\rangle\\ \nn
&=& A^2\alpha^2\intinf |\tilde\psi(\nu)|^2|\tilde r(\nu)|^2\de \nu+ \langle \hat Q_e^2\rangle,
\end{eqnarray}
where the variance of the electronic noise component is given by
\begin{eqnarray}
\langle \hat Q_e^2\rangle
&=& \intinf\intinf \langle \hat i_e(t)\hat i_e(t+\tau)\rangle\psi(t)\psi(t+\tau)\de t\de\tau\nn \\
&=& \intinf S_e(\nu)|\tilde\psi(\nu)|^2\de\nu.\label{QeCV}
\end{eqnarray}
Combining the above two results with Eq.~\eqref{SnuCV}, we find
\begin{equation}\label{QmeasCVvar1}
\langle \hat Q_{\rm meas}^2\rangle = \intinf S(\nu) |\tilde\psi(\nu)|^2 \de\nu.
\end{equation}
Equations \eqref{QeCV} and \eqref{QmeasCVvar1} lead us to an important conclusion: by knowing the homodyne output spectra $S(\nu)$ and $S_e(\nu)$, as well as the weight function $\psi(t)$, one can predict the fraction of the electronic noise in the measured quadrature variance in an arbitrary temporal mode. This, as discussed above, directly translates into an equivalent optical loss.
In the case of a high-bandwidth detector, when $S(\nu)$ and $S_e(\nu)$ can be assumed constant over the support of $\tilde\psi(\nu)$, we have
\begin{equation}\label{}
1-\eta_e\approx \left[\frac {S_e(\nu)}{S(\nu)}\right]_{\textrm{signal bandwidth}}.
\end{equation}
This quantity, which we call \emph{clearance} of the detector's shot noise over the electronic noise, is one of the primary characteristics of any HD circuit.
\subsection{Pulsed regime}
Now let us suppose that the LO is pulsed, with the pulse width much shorter than the time resolution of the electronics. In this case, \eeqref{cpracHD} takes the form
\begin{equation}\label{pracHD}
\hat i(t)= A\alpha_p\hat Q r(t)+\hat i_e(t),
\end{equation}
where the pulse is assumed to occur at $t=0$, $\hat Q=\intinf \alpha(t) \hat q(t)\de t/\alpha_p$ is the normalized quadrature operator corresponding to the signal mode defined by the shape of the LO pulse, with
$\alpha_p=\sqrt{\intinf|\alpha(t)|^2\de t}$ being the effective amplitude of the local oscillator pulse. In other words, neglecting the electronic noise, the shape of the HD response to a single short pulse is given by the detector's response function.
The quadrature measurement is obtained by integrating the homodyne photocurrent over a certain time interval:
\begin{equation}\label{Qmeas}
\hat Q_{\rm meas}=\int\limits_{t_1}^{t_2}\hat i(t)\de t=A\alpha_p \hat Q\int\limits_{t_1}^{t_2}r(t)\de t+\hat Q_e,
\end{equation}
with $$\hat Q_e=\int\limits_{t_1}^{t_2}\hat i_e(t)\de t.$$ The optimal choice of the integration limits is determined by the bandwidths of the detector's electronic noise and the temporal width of its response function.
When the local oscillator is a train of pulses with repetition period $T$, the HD output current is given by
\begin{equation}\label{pracHDrep}
\hat i(t)= A\alpha_p\sum\limits_{j=-\infty}^{\infty}\hat Q_j r(t-jT)+\hat i_e(t),
\end{equation}
where $\hat Q_j$ is the quadrature operator of the $j$th pulsed signal mode, with the pulse of interest having index $j=0$. If the response function is nonzero over an interval longer than $T$, the quadrature measurement is contaminated by that of the neighboring pulses:
\begin{equation}\label{QmeasSum}
\hat Q_{{\rm meas},0}=A\alpha_p \sum\limits_{j=-\infty}^{\infty}R_j\hat Q_j +\hat Q_e, \end{equation}
where $R_j=\int\limits_{t_1}^{t_2}r(t-jT)\de t$. The sum in \eeqref{QmeasSum} defines a new measured mode whose state is not necessarily identical to that in the $j=0$th pulsed mode. The corresponding mode matching efficiency (neglecting the electronic noise) is given by
\begin{equation}\label{}
\eta_b=\frac{R_0^2}{\sum\limits_{j=-\infty}^{\infty}R_j^2}.
\end{equation}
This efficiency is plotted in Fig.~\ref{pulsedefftheory} for the response function of Gaussian shape
as a function of the 3-dB bandwidth of the detector response function spectrum. As we see, a detector bandwidth of at least $~0.4/T$ is required for $\eta_b$ to exceed 99\%.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.8\columnwidth]{pulsedefftheory.eps}
\caption{Effective efficiency of the HD associated with temporal overlap of responses to different pulsed modes, as a function of the 3-dB bandwidth of the electronics. The response function is assumed Gaussian. The solid line corresponds to a short integration interval $[-\epsilon,+\epsilon]$; the dashed line to the integration interval of length $T$. Both integration intervals are centered at the peak of the response function. \label{pulsedefftheory}}
\end{center}
\end{figure}
If $r(t)$ is known, so are all $R_j$ and the effect of finite bandwidth can be reversed by means of discrete deconvolution, akin to the continuous case. However, partial reversal can be implemented even if the response function is not known, provided that $\eta_b$ is sufficiently high, i.e. $|R_j|\ll|R_0|$ for $j\ne 0$, as follows. In a typical detector, $R_j$ are negligibly small for $j>0$: most of the ringings occur \emph{after} the optical pulse that generates the response. This is the situation, for example, with the detector assembled in this work
Then we have, according to \eeqref{QmeasSum}, and because $\langle\hat Q_j\hat Q_k\rangle=\delta_{jk}$ in the vacuum state,
\begin{equation}\label{}
\langle \hat Q_{\rm meas,0}\hat Q_{{\rm meas},i}\rangle=A^2\alpha^2_p\sum\limits_j R_jR_{j-i}\approx A^2\alpha^2_p R_0R_{-i},
\end{equation}
where $\hat Q_{{\rm meas},i}$ denotes the quadrature measurement for the $i$th pulse. The above correlation can be easily obtained experimentally, from which one can determine
\begin{equation}\label{}
\frac {R_{-i}}{R_0}\approx \frac{\langle \hat Q_{\rm meas,0}\hat Q_{{\rm meas},i}\rangle}{\langle \hat Q_{\rm meas,0}\hat Q_{{\rm meas},0}\rangle}.
\end{equation}
One then calculates
\begin{equation}\label{}
Q'_{{\rm meas},0}=Q_{{\rm meas},0}- \sum\limits_{i=1}^{\infty}\frac{\langle \hat Q_{\rm meas,0}\hat Q_{{\rm meas},i}\rangle}{\langle \hat Q_{\rm meas,0}\hat Q_{{\rm meas},0}\rangle} Q_{{\rm meas},i}
\end{equation}
for each experimentally measured quadrature, thereby eliminating contamination from neighboring pulses. The resulting quadrature values are then renormalized and used in quantum state reconstruction. By means of this technique, the reconstruction efficiency has been improved in Ref.~\cite{hui:09}.
We now use \eeqref{Snu} to determine the spectral power of the HD output in the pulsed regime.
The ensemble average of $i(t)i(t+\tau)$ is a periodic function of time $t$, hence we can write
\begin{eqnarray}\label{}
\langle \hat i(t)\hat i(t+\tau)\rangle&=&\frac 1T (A\alpha_p)^2\\ \nn &&\hspace{-2cm}\times\sum\limits_{j,k=-\infty}^{\infty}\langle\hat Q_j\hat Q_k\rangle\int\limits_{-T/2}^{T/2}r(t-jT)r(t+\tau-kT)\de t\\ \nn&&\hspace{-2cm}+\langle \hat i_e(t) \hat i_e(t+\tau)\rangle_t.
\end{eqnarray}
In the vacuum state,
\begin{equation}\label{}
\langle \hat i(t)\hat i(t+\tau)\rangle=\frac 1T (A\alpha_p)^2\intinf r(t)r(t+\tau)\de t+\langle \hat i_e(t) \hat i_e(t+\tau)\rangle_t.
\end{equation}
Accordingly,
\begin{equation}\label{Sm}
S(\nu)=\frac 1T (A\alpha_p)^2 |\tilde r(\nu)|^2+S_e(\nu).
\end{equation}
We see that in spite of the pulsed character of the local oscillator, the HD spectrum is determined by the Fourier transform of its response function akin to the continuous case. An important difference is the multiplication by the pulse repetition rate: when the separation between the pulses is increased, the spectral power reduces proportionally.
In contrast to the continuous regime, in the pulsed case the evaluation of the equivalent efficiency \eqref{etae} requires knowledge of $r(t)$; information on the spectra $S(\nu)$ and $S_e(\nu)$ is not sufficient. Let us, however, consider a practically important particular case when the bandwidth of both the detector's response and the electronic noise greatly exceed the laser repetition rate. Suppose the integration in \eeqref{Qmeas} is done over the time interval $[-T_0/2,T_0/2]$ with $T_0<T$. Then we have
\begin{equation}\label{Qmeas2}
\langle \hat Q_{\rm meas}^2 \rangle=\left\langle \left(A\alpha_p\hat Q\int\limits_{-T_0/2}^{T_0/2}r(t)\de t\right)^2\right\rangle+\langle \hat Q_e^2\rangle.
\end{equation}
We assume that the temporal width of the function $r(t)$ is much less than $T_0$, so the integration limits can be replaced by $\pm\infty$. We then have
\begin{equation}\label{Qmeas2hb}
\langle \hat Q_{\rm meas}^2 \rangle=(A\alpha_p)^2|\tilde r(0)|^2+\langle \hat Q_e^2\rangle,
\end{equation}
where
\begin{eqnarray}\label{Qe2hb}
Q_e &=& \left\langle \left(\int\limits_{-T_0/2}^{T_0/2}\hat i_e(t)\de t\right)^2\right\rangle \\ \nn
&=& \left\langle \int\limits_{-T_0/2}^{T_0/2}\int\limits_{-T_0/2}^{T_0/2}\hat i_e(t_1)\hat i_e(t_2)\de t_1\de t_2\right\rangle\\ \nn
&=& \int\limits_{-T_0/2}^{T_0/2}\int\limits_{-T_0/2-t_1}^{T_0/2-t_1}\langle \hat i_e(t_1)\hat i_e(t_1+\tau)\rangle\de \tau\de t_1\\ \nn
&\approx& \int\limits_{-T_0/2}^{T_0/2}\int\limits_{-\infty}^{+\infty}\langle \hat i_e(t_1)\hat i_e(t_1+\tau)\rangle\de \tau\de t_1\\ \nn
&=& T_0 S_e(0).
\end{eqnarray}
Here we again took advantage of the high bandwidth of the electronic noise to modify the integration limits. We also used the fact that the electronic noise $i_e(t)$ is a stationary stochastic process to eliminate the dependence on $t_1$. Substituting Eqs.~\eqref{Sm}--\eqref{Qe2hb} into \eeqref{etae}, we find
\begin{equation}\label{etaeS}
\eta_e=1-\frac {T_0} T \frac{\hat S_e(0)}{ S(0)}.
\end{equation}
In other words, the shot-to-electronic-noise clearance measured in the pulsed regime yields a too conservative estimate for the equivalent loss associated with the electronic noise in the high-bandwidth limit. In order to minimize this loss, one needs to choose the integration interval that is as short as possible, but still accommodates the entire detector response function.
However, if the detector bandwidth does not greatly exceed the laser repetition rate, we have $T_0\simeq T$ and the factor of $T_0/T$ in the above equation can be neglected. The exact value for the equivalent loss in this case cannot be determined from the spectra because it depends on the shape of the detector's response function $r(t)$.
\section{Design and challenges}\label{designSec}
Now we present an efficient wideband HD constructed with easily available components, as well as tips to solve the most common challenges found when building such a device. We show the results given by our HD in experiments with a pulsed laser source (76 MHz repetition rate) as well as with a continuous wave laser. These features give our HD a great versatility for applications in different kinds of experiments.
\begin{figure*}[t]
\begin{center}
\includegraphics[scale=0.65]{schm1.eps}
\caption{HD Circuit diagram\label{fig:schm}}
\end{center}
\end{figure*}
Our electronic cicruit is shown in Fig.~\ref{fig:schm}. The values and identification names are as noted on the schematic. Provision has been made for a second stage of voltage gain which, however, proved to be unnecessary.
We use a homemade $\pm 15$ V DC power supply. To avoid possible leakage of any ambient noise to the HD through the voltage supply lines, each of them is filtered for both high and low frequencies (components L1, L2, C13, C14, C15, C16 that include a ferrite choke, a 100 $\mu$f electrolytic and a 0.1 $\mu$f ceramic capacitors). The use of long power supply lines in the circuit can create parasitic inductances making the circuit susceptible to instabilities. Because it is not possible to make all the tracks short, we perform the filtering in two stages, at the beginning and the end of each power supply track.
The filtered $\pm 15$ Volt supply rails are each further separated into 3 isolated lines by incorporating 3 networks consisting of a fast Schottky diode (for improving the stability of the power supplies), a tantalum 6.8 $\mu$f capacitor and a 0.1 $\mu$f ceramic capacitor (components C1--C12 and D3--D8). These isolated supplies in turn feed $\pm 12$ V regulators for the photodiodes (IC3 and IC8) and two pairs of $\pm 5$ V regulators for the first and second stage operational amplifiers (IC2, IC5, IC4 and IC7). Care has been taken to isolate and regulate the supply voltage for the detectors and amplifiers in order to provide as stable and oscillation free platform as possible. Bypass 0.1 $\mu$f capacitors are located at the regulator outputs very close to the amplifiers (C18--C23).
One of the primary challenges in achieving a high-bandwidth HD is associated with the photodiodes. Their P-N junction has a terminal capacitance which has to be significant for high quantum efficiency of detection. The feedback resistance of the amplifier will form a low pass filter with the terminal capacitance, thus limiting the high-frequency response of the circuit. Furthermore, this capacitance combined with other capacitances and inductances in the circuit board can give rise to instabilities and oscillations associated with a low pass filter configuration.
The terminal capacitance of the photodiodes can be reduced by increasing the reverse bias voltage across them. This, however, increases the dark current that contributes to the electronic noise \cite{zav:02}. Consequently, special care must be taken for choosing photodiodes with very low capacitance and dark current, as well as for a printed circuit board (PCB) design that avoids or compensates any capacitance that could produce oscillations (instabilities) in the frequency response of the HD.
We have chosen to use Hamamatsu S5972 photodiodes. This photodiode has a 91\% quantum efficiency at a wavelength of 780 nm and a high cut-off frequency (500 MHz) when supplied in a 12 V reverse voltage configuration. Other reasons for this choice include a very low dark current (11 pA at 12V reverse voltage) and a low terminal capacitance (2.8 pf at 12V reverse voltage).
The amplification circuit starts at the junction of the two S5972 photodiodes, which is the differential sum point for the input to the first stage amplifier. The amplification of the difference signal is carried out by a single Texas Instruments OPA847 operational amplifier (op-amp) in trans-impedance configuration with a trans-impedance gain of 4 k$\Omega$ (IC1). For testing purposes we place a jumper (J1) in the inverting input of the op-amp. In order to have control of the DC offset of output signal produced by the the op-amp we place in its non-inverting input a capacitor of 0.1 $\mu$f in parallel with a variable resistor of 10 k$\Omega$ both connected to ground (C17 and R4). With the help of the variable resistor we can maintain the DC offset at the zero reference point to avoid saturation in the op-amp. The optional second stage is designed to use an OPA847 in inverting amplifier configuration. J2 and J3 are the positions of its input and feedback resistances.
An important component of the circuit is a custom made variable capacitor (C24) placed in parallel to the feedback resistance R2. This capacitor is constructed by twisting two wires with a variable number of twists. By twisting (or untwisting) these wires we can vary their capacitance; although it is very small, it is enough to change the response of the circuit. In our case it is used to flatten the spectral response of the HD.
Finally, resistors R3, R5 and R6 are used for impedance matching of the output signal of the amplifier with a 50$\Omega$ coaxial connector.
All components are soldered on a specially designed printed circuit board (Fig.~\ref{fig:pcb}). When designing the layout for the circuit we take the following factors into consideration.
\begin{itemize}
\item The tracks on the PCB are kept as short as possible to avoid parasitic inductances.
\item A ground (GND) plane, essential for high frequency circuits, and a double-sided board design are used. The back side of the PCB is mainly reserved for the GND plane, with the fewest possible discontinuities. This separation between tracks and components in one side and GND on the other minimize possible capacitances between the tracks and the ground plane of the circuit.
\item The use of surface mount components is preferred for the high frequency regime of electronics.
\item In order to provide a mechanically stable platform for the HD, the circuit module is mounted on a 0.25-inch aluminum plate.
\item To ensure a maximally balanced response, we designed our circuit to be as symmetrical as possible.
\item The HD is shielded by a custom-made metal box to avoid any environmental noise impact on the circuit.
\item To minimize the parasitic capacitance the two photodiodes were placed closely together in the upper left corner of the board.
\item Due to its relatively large size, the offset adjustment potentiometer (R4) for the first stage is placed at the bottom of the printed circuit board to avoid the issue of having long tracks on the upper side.
\end{itemize}
\begin{figure*}[t]
\begin{center}
\includegraphics[width=6in]{pcb.eps}
\caption{HD Layout: (Left) Bottom layer where the ground plane is; (Right) Upper layer, connection tracks and OPAmps\label{fig:pcb}}
\end{center}
\end{figure*}
\begin{figure}[h]
\begin{center}
\includegraphics[width=2.5in]{setup.eps}
\caption{Balanced HD experimental set-up\label{fig:setup}}
\end{center}
\end{figure}
\section{Performance and characterization}
\subsection{Pulsed regime}\label{pusedSec}
After having built the HD we proceed to test its performance. To that end, we place it in the experimental set-up shown in Fig.~\ref{fig:setup}. Our LO is obtained from a mode-locked Coherent MIRA 900 Ti:Sapphire laser producing 1.8 ps pulses with a repetition rate of 76 MHz, central wavelength of 791 nm. The HD beam splitter configuration is implemented by using a half wave plate and a polarizing beam splitter. The two beams coming out of the beam splitter are focused into the photodiodes.
The HD is balanced by first adjusting the waveplate in order to equalize the responses of the photodiodes, compensating for a possible difference in their quantum efficiencies, thereby minimizing the spurious signal at the local oscillator repetition rate. A further crucial step in balancing the HD is to equalize the path lengths of the two beams entering the photodiodes by using a XY translation stage on which the HD is mounted. This is necessary because the pulses, if arriving at the photodiodes at different times, can lead to subtraction pulse having a bipolar shape. Even after these alignment steps, the subtraction may not be perfect due to different capacitances of the two photodiodes.
Subsequently, the HD is tested for electronic instabilities and oscillations. This is done by observing the output signal of the HD (with the LO power of 6 mW) with a spectrum analyzer in a range from 100 kHz up to 3 GHz. If present, instabilities produce peaks in the spectral response at frequencies different from the repetition rate of the LO and its harmonics. They can be removed by adjusting the values of resistances R2, R3, R5 and R6 as well as the custom made variable capacitor C24. Additional instability sources can be associated with the ambient noise or leakage through the power supply lines; in this case, the elements of the power supply filtering sections must be changed.
In the final step of adjustments we flatten the spectral response of the HD. This is done by changing the value of the capacitance C24 while observing the spectrum of the HD output. Fig.~\ref{fig:spectralresponsepulse} (dashed and dotted lines) shows the effect of this adjustment in the spectral profile of the HD.
In order to characterize the detector, we first measure the shot noise clearance. To this end, we increase the LO power to 12 mW. The choice of this power level is determined by the need to achieve the highest shot-to-electronic noise clearance, while at the same time ensuring stable and saturation-free operation of the HD. With the given wavelength and repetition rate, this LO power corresponds to $6.2 \times 10^{8}$ photons per pulse, and will produce $5.6 \times 10^{8}$ photoelectrons per pulse, which corresponds to a shot noise of $24,000$ electrons per pulse. In Fig.~\ref{fig:spectralresponsepulse} (dotted vs. dash-dotted lines) we show the clearance between electronic noise and shot noise, which has a value of 13dB, correponding to $\eta_e=0.95$. The 3-dB bandwidth of the HD is 80 MHz, which implies that it can be applied, without significant efficiency loss, to pulsed lasers with repetition rates up to 250 MHz (Fig.~\ref{pulsedefftheory}).
We verify that the power of the observed HD output signal grows linearly with the LO power, which is the signature of the shot noise \cite{bac:04,lvov:09}. In Fig.~\ref{fig:linear} we show measurements of the shot noise level for five different frequencies depending on the LO power. For frequencies up to 100 MHz, the HD behaves linearly, although the slope is reduced for higher frequencies.
\begin{figure}[t]
\begin{center}
{
\includegraphics[width=0.9\columnwidth]{spectralresponsepulse.eps}
}
\caption{Spectral response of the HD at 12 mW of LO power. The large peaks are the laser repetition rate and its second harmonic. \emph{Inset}: The response at 100 $\mu$W of LO power, necessary to calculate the CMRR. Grey (top) trace: one photodiode blocked; Black (bottom) trace: both photodiodes illuminated\label{fig:spectralresponsepulse}}
\end{center}
\end{figure}
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.8\columnwidth]{linear.eps}
\caption{Response of the HD respect to the LO power for different fixed frequencies. The linear fits are also shown. Red dash line (plus): 80MHz; Green dot line (diamond): 60MHz; Black continous line (circle): 40MHz; Blue dot dash line (cross): 100MHz.\label{fig:linear}}
\end{center}
\end{figure}
In order to determine the CMRR value, we measure the spectral response of the HD at the repetition frequency of the LO at a power of 100 $\mu$W in two cases: when both photodiodes are illuminated, and when only one photodiode is illuminated (inset of Fig.~\ref{fig:spectralresponsepulse}). The measurement is performed with a low LO power to avoid saturation of the HD in the absence of subtraction. We find that our HD has a CMRR of 52.4 dB, which corresponds to a reduction by a factor of $2.4\times10^{-3}$ in the photocurrent corresponding to the common mode, yielding $6.7\times 10^5$ residual photoelectrons per pulse due to imperfect subtraction \cite{chi:11}, which is much greater than the average number of photoelectrons per pulse due to the shot noise. This is the reason why, even with the optimized alignment, the pulsed HD spectrum exhibits a peak at the pulse repetition rate.
Theoretically, this residual periodic signal leads to a constant displacement in the quadrature measurement, which is easy to average out. In practice, however, the signal tends to spontaneously change its magnitude every few seconds. Furthermore, the presence of that signal complicates computer acquisition of the HD output because of the limited digitizer resolution. Therefore we use two notch filters (MFC 6367) to eliminate the output components at 76 and 152 MHz.
The HD output signal is further processed by a 100 kHz high pass filter to remove the DC offset. To clean the signal and to avoid any noise at higher frequencies, two low pass filters with a 100 MHz cut-off are used. The resulting HD spectrum is shown in Fig.~\ref{fig:spectralresponsepulse} (solid line).
\subsection{Continuous regime}
In order to check the versatility of our scheme, we analyze the performance of another HD, constructed similarly to the first one, in the continuous wave regime by using a TekhnoScan Ti:Sapphire laser with a center wavelength of 795 nm and a bandwidth of 5 kHz. We use the same experimental set-up as in Fig.~\ref{fig:setup}. We analyze the response of the HD in the time domain by acquiring 30 traces of 100-$\mu$s duration, each containing $2 \times 10^{5}$ points (sampling frequency of 2GHz). Thereafter, we compute the autocorrelation function for each trace, and then average it over the 30 traces [Fig.~\ref{fig:spectralresponsecont}(a)].
\begin{figure}[h]
\begin{center}
{
\includegraphics[width=0.9\columnwidth]{cvspectrum.eps}
}
\caption{characterization of the HD in the continuous regime. a) Autocorrelation function. b) Electronic spectrum of the shot noise obtained with a spectrum analyzer (blue trace --- top); the black (bottom) trace is the electronic noise, and by Fourier transform of the autocorrelation function (red trace --- middle). The vertical position on the logarithmic scale of the red trace with respect to the spectrum analyzer traces is chosen arbitrarily. c) Effective efficiency \eqref{etae} associated with the electronic noise as a function of the signal wavefunction bandwidth (in units of the inverse FWHM of its temporal mode). \label{fig:spectralresponsecont}}
\end{center}
\end{figure}
We then acquire the HD output spectrum, both in the presence and absence of LO, using a spectrum analyzer. As evidenced by Fig.~\ref{fig:spectralresponsecont}(b), the shot noise spectrum is similar, up to a constant factor, to that obtained by Fourier transform to the autocorrelation function in agreement with the Wiener-Khintchine theorem. We find this detector to exhibit a 3-dB bandwidth of 100 MHz and the shot-to-electronic noise clearance ranging between 10 and 18 dB.
In Fig.~\ref{fig:spectralresponsecont}(c) we plot the effective loss \eqref{etae} associated with the electronic noise under the assumption that the detector is used to measure field quadratures in a Gaussian temporal mode. The loss is calculated according to Eqs.~\eqref{QeCV}, and \eqref{QmeasCVvar1} given the measured spectra [Fig.~\ref{fig:spectralresponsecont}(b)]. We find that the effective loss strongly varies dependent on the shape and width of the temporal mode.
Table \ref{tab:table1} shows the features of our device in comparison with other HDs reported in the literature. As we see, our detector compares favorably with its counterparts: it shows a unique combination of the bandwidth, CMRR and the electronic noise clearance.
\begin{table}[h]
\caption{\label{tab:table1}Comparison between various HD's.}
\begin{ruledtabular}
\begin{tabular}{l|r|r|r|r|r|r|r}
Characteristics&Ours&\cite{han:01}&\cite{chi:11}&\cite{zav:02}&\cite{nie:09}&\cite{oku:08}&\cite{had:09}\\
\hline
Wavelength (nm) &791 &790 &1550 &786 &860 &1064 &--\\
3 dB bandwidth (MHz)&100 &1 &100 &$\sim$50 &100 &250 &54\\
CMRR (dB) &52.4 &85 &46 &42 &-- &45 &61.8\\
Clearance (dB) &13 &14 &13 &$\sim$5 &10 &$\sim$7.5 &12\\
\end{tabular}
\end{ruledtabular}
\end{table}
\section{Summary}
We presented a comprehensive theory and a detailed recipe for designing, building and troubleshooting a wideband balanced detector for highly accurate time-domain quantum measurements. We showed that by following these recommendations a homodyne detector with a shot-noise clearance of 13 dB at 12 mW of local oscillator power, a CMRR value of 52.4dB, a flat response up to 100 MHz can be constructed in an easy way and by only using a single trans-impedance amplification stage. We also have shown that this HD can be applied to both pulsed and CW configurations.
We thank Frank Vewinger, Nitin Jain and Ryan Thomas for helpful discussions. This work has been supported by NSERC, CIFAR, Quantum\emph{Works} and ARC (Grant DP1094650).
\bibliographystyle{aip}
\input{hd-rsi.bbl}
\end{document}
|
\section{Introduction}
In the past few years one of the most important experimental
progresses in the study of the charmed meson spectrum is the
establishment of the lowest $P$-wave charmed-strange mesons, i.e.
$D_{s0}(2317)$, $D_{s1}(2460)$, $D_{s1}(2536)$, and $D_{s2}(2573)$
as now listed in Particle Data Group (PDG) 2010
Edition~\cite{Nakamura:2010zzi}. Since the first observation by
BaBar Collaboration~\cite{Aubert:2003fg}, the spin-0 state
$D_{s0}(2317)$ and spin-1 $D_{s1}(2460)$ (later confirmed by
Belle~\cite{Krokovny:2003zq} and CLEO~\cite{Besson:2003cp}) have
initiated tremendous interests in its property and internal
structure. These two states have masses lower than the potential
model predictions, and their widths are rather narrow. It is somehow
agreed that their low masses are caused by the open $D K$ and $D^*K$
thresholds, respectively, and as a consequence, their narrow decay
widths are due to the dominant isospin-violating decays, i.e.
$D_{s0}(2317)\to D_s\pi$ and $D_{s1}(2460)\to D_s^*\pi$ (see the
review of Refs.~\cite{Swanson:2006st,Godfrey:2005ww} and references
therein).
The heavy-light $Q\bar{q}$ system is an ideal platform for testing
the internal constituent quark degrees of freedom. In the heavy
quark limit the heavy quark spin is conserved and decoupled from the
light quark degrees of freedom, which are characterized by the total
angular momentum ${\bf j}_q\equiv {\bf s}_q+{\bf L}$, where ${\bf
s}_q$ is the light quark spin and ${\bf L}$ is its orbital angular
momentum. With $j_q=1/2$ and $j_q=3/2$, one can arrange those four
$P$-wave states into two classes, i.e. $J^P=0^+, \ 1^+$ and
$J^P=1^+, \ 2^+$, respectively, where $J$ is the meson spin as a sum
of the heavy quark spin ${\bf S}_Q$ and ${\bf j}_q$. For the axial
vector states in the charmed and charmed-strange meson spectrum,
since they are not charge conjugation eigenstates, state mixings
between the $^3P_1$ and $^1P_1$ configurations are allowed. In the
case of charmed and charmed-strange heavy-light system when the
heavy quark symmetry is broken at order of $1/m_c$, it would be
interesting to study the mechanism that causes deviations from the
ideal mixing scenario, i.e. breakdown of the heavy quark symmetry.
This forms our motivation in this work. As mentioned earlier,
$D_{s1}(2460)$ and $D_{s1}(2536)$ lie near the threshold of $D^* K$
and both couple to $D^* K$ strongly via a relative $S$ wave. It
gives rise to coupled channel effects in the mass shifts of
potential quark model calculations in comparison with the observed
values~\cite{Simonov:2004ar,Badalian:2007yr,Coito:2011qn}, and
produces state mixings between the $^3P_1$ and $^1P_1$
configurations. Similar mechanism has been studied in the
$a_0(980)$-$f_0(980)$ mixing in Ref.~\cite{Wu:2007jh}. Determination
of the mixing angle should be useful for understanding the property
and internal structure of these two axial vector states.
We mention that various solutions have been proposed in the
literature to explain the observed results for $D_{s1}(2460)$ and
$D_{s1}(2536)$. For instance, $D^*K$ molecule or tetra-quark
configuration have been investigated in
Refs.~\cite{Close:2005se,Barnes:2003dj,Faessler:2007us}. In
Ref.~\cite{Guo:2006rp}, $D_{s1}(2460)$ is explained as a dynamically
generated state. The mixing angle has also been calculated in the
quark model~\cite{Godfrey:1985xj,Yamada:2005nu} but with large
uncertainties from the quark spin-orbital interactions. In this
work, we investigate the two-state mixing propagator matrix which
respects the unitarity constraint in a chiral quark model. We will
show that the coupled channel effects via intermediate hadron loops
can provide a simultaneous determination of the masses, widths and
mixing angles of these two axial vector states. We also mention that
the coupled channel effects on the ${}^3P_1$ and ${}^1P_1$ mixing
was recently studied in Ref.~\cite{Zhou:2011sp}, where the the
couplings were extracted in the ${}^3P_0$ model and a subtracted
dispersion relation was applied to evaluate the hadron loops. In our
approach we use the chiral quark model to extract the couplings and
vertex form factor. We then extend the quark model form factor to a
covariant form which can be applied on a general ground to much
broader cases.
The paper is organized as follows. In Sec.~\ref{sec:cce}, we give
the basic formulas of two-state mixings via coupled channel
propagators. In Sec.~\ref{sec:vertices}, the relevant coupling form
factors are determined by the chiral quark model. In
Sec.~\ref{sec:propagator} the propagator matrix is calculated in
detail. Section~\ref{sec:mass} is devoted to show our numerical
results for the mass and mixing parameters. The experimental
constraints for the mixing angle are presented in
Sec.~\ref{sec:exp}. A summary is given in the last Section. In
Appendix~\ref{app:1} the detailed definition and calculation of a
special function used in the evaluation of the loop integrals with
exponential form factors are provided.
\section{Mixing through coupled channel effect}\label{sec:cce}
We use $|a\rangle$ and $|b\rangle$ to present two pure states in the
quark model. If they can couple to common final states, there will
be a transition between them via single particle irreducible (1PI)
diagrams as shown in Fig.~\ref{chain}.
\begin{figure}[htbp]
\includegraphics[scale=0.8]{fig1}
\vspace{-17.5cm}
\caption{Transition through intermediate states}\label{chain}
\end{figure}
The propagator matrix of $|a \rangle$ and $|b\rangle$ can be expressed as
\begin{equation}\label{eqn1}
G_{ab}=\left(\begin{array}{c}
\langle a |\\
\langle b |
\end{array}
\right)
\hat{S}
\left( |a\rangle , |b\rangle
\right) \ .
\end{equation}
The physical states $|A\rangle$ and $|B\rangle$ should be a mixture of $|a \rangle$ and $|b\rangle$,
\begin{equation}\label{eqn2}
\left(\begin{array}{c}
|A\rangle \\
|B\rangle
\end{array}
\right)
=
\left(
\begin{array}{cc}
\cos \theta & - \sin \theta e^{i \phi}\\
\sin \theta e^{-i \phi} & \cos \theta
\end{array}
\right)
\left(\begin{array}{c}
|a\rangle \\
|b\rangle
\end{array}
\right)
=
R(\theta,\phi)
\left(\begin{array}{c}
|a\rangle \\
|b\rangle
\end{array}
\right)
\end{equation}
where $R(\theta,\phi)$ is the mixing matrix, $\theta$ is the mixing
angle, and $\phi$ is a possible relative phase between $|a \rangle$
and $|b\rangle$. Then the propagator matrix of $|A\rangle$ and
$|B\rangle$ is
\begin{equation}\label{eqn3}
G_{AB}=R G_{ab} R^{\dagger} \ .
\end{equation}
The physical propagator matrix $G_{AB}$ should be a diagonal matrix.
Thus, we can determine the mixing parameters $\{\theta,\phi\}$ by
diagonalizing the propagator matrix $G_{ab}$.
In the present case, we set $|a\rangle=| {}^3P_1 \rangle$,
$|b\rangle=| {}^1P_1 \rangle$, $|A\rangle=|D_{s1}(2460)\rangle$ and
$|B\rangle=|D_{s1}(2536) \rangle$ as in Ref.~\cite{Badalian:2007yr}.
The mixing scheme is
\begin{eqnarray}\label{eqn2:1}
|D_{s1}(2460)\rangle &=& \cos\theta | {}^3P_1 \rangle - \sin\theta e^{i \phi} | {}^1P_1 \rangle \nonumber \\
|D_{s1}(2536)\rangle &=& \sin\theta e^{-i\phi} | {}^3P_1 \rangle + \cos\theta | {}^1P_1
\rangle \ ,
\end{eqnarray}
where states $| {}^3P_1 \rangle$ and $| {}^1P_1 \rangle$ can be
rotated to the eigenstates in the heavy quark limit:
\begin{equation}\label{eqn2:2}
\left(\begin{array}{c}
| {}^3P_1 \rangle \\ |{}^1P_1 \rangle
\end{array}\right)=
\left(\begin{array}{cc}
\sqrt{\frac{2}{3}} & \sqrt{\frac{1}{3}} \\
-\sqrt{\frac{1}{3}} & \sqrt{\frac{2}{3}}
\end{array}\right)
\left(\begin{array}{c}
| j=\frac{1}{2} \rangle \\ | j=\frac{3}{2} \rangle
\end{array}\right) \ .
\end{equation}
The mixing angle $\theta$ defined in Eq.~(\ref{eqn2:1}) can be
related to $\theta'$ defined in $j=1/2$ and $j=3/2$ bases:
\begin{equation}\label{eqn2:3}
\theta=\theta'+35.26^\circ \ .
\end{equation}
Considering parity conservation, the important intermediate states
that can couple to $D_{s1}(2460)$ and $D_{s1}(2536)$ are $D^* K$,
$D_{s}^{*} \eta$ and $D K^*$, of which the thresholds are listed in
Table~\ref{tab1}.
\begin{table}[htbp]
\centering
\caption{The thresholds of intermediate states for $D_{s1}(2460)$ and $D_{s1}(2536)$.}\label{tab1}
\begin{tabular}{c|c|c|c|c|c}
\hline\hline
Intermediate states & $D^{*0} K^+$ & $D^{*+} K^0$ & $D_{s}^{*} \eta$ & $D^0 K^{*+}$ & $D^+ K^{*0}$\\
\hline
Threshold (GeV) & 2.501 & 2.508 & 2.660 & 2.756 & 2.761 \\
\hline\hline
\end{tabular}
\end{table}
If all the particles involved are scalars or pseudoscalars,
Fig.~\ref{chain} will only represent sums of infinite geometric series
and the resulting propagator matrix $G$ becomes \cite{Wu:2007jh}
\begin{equation}\label{eqn4}
G_{ab}=\frac{1}{D_a D_b-D_{ab}^2} \left(\begin{array}{cc}
D_b & D_{ab} \\
D_{ba} & D_a
\end{array}
\right) \ ,
\end{equation}
where $D_a$ and $D_b$ are the denominators of the single propagators
of $|a\rangle$ and $|b\rangle$, respectively, and the mixing term
$D_{ab}$ is the sum of all 1PI diagrams, which satisfies
$D_{ba}$=$D_{ab}$. But from Table \ref{tab1}, we find that the
particles involved in the present case can be scalars, vectors or
axial-vectors. There are five diagrams for the mixing of
$D_{s1}(2460)$ and $D_{s1}(2536)$ as shown in Fig. \ref{mixterm}.
\begin{figure}[htbp]
\includegraphics[scale=0.7]{fig2}\\
\vspace{-14cm}
\caption{Mixing term for $D_{s1}(2460)$ and $D_{s1}(2536)$.}\label{mixterm}
\end{figure}
The mixing term can be generally divided into transverse and
longitudinal terms:
\begin{equation}\label{eqn5}
D_{ab}^{\mu \nu}\equiv \Pi_{ab} P^{\mu \nu} + B_{ab} Q^{\mu \nu} \ ,
\end{equation}
where $P^{\mu \nu}\equiv g^{\mu \nu}-p^{\mu}p^{\nu}/p^2$ and $Q^{\mu
\nu}\equiv p^{\mu}p^{\nu}/p^2$ are the transverse and longitudinal
projector, respectively, and satisfy
\begin{equation}\label{eqn5a}
P^{\mu \nu}P_\nu^{\lambda}=P^{\mu \lambda}, \
Q^{\mu \nu}Q_\nu^{\lambda}=Q^{\mu \lambda}, \
P^{\mu \nu}Q_\nu^{\lambda}=0 \ .
\end{equation}
Next we concentrate on the evaluation of the propagator matrix
$G^{\mu \nu}$ for axial vector states. The numerator of the vector
propagator is $g^{\mu \nu}-p^{\mu}p^{\nu}/m^2$ and can be generally
expressed as $P^{\mu \nu}+ \Delta Q^{\mu \nu}$ where
$\Delta=1-p^2/m^2$. With the properties of Eq.(\ref{eqn5a}), the
geometric sums, e.g. $\langle a | \hat{S} | b\rangle$ in
Fig.~\ref{chain}, can be taken for the transverse and longitudinal
terms independently. After include the self-energy functions
$\Pi_a^{\mu \nu}$ and $\Pi_b^{\mu \nu}$, the complete propagator
matrix for the $1{}^3 P_1$ and $1{}^1 P_1$ states becomes
\begin{equation}\label{eqn6}
G_{ab}^{\mu \nu}=iP^{\mu \nu}
\frac{\bar{G}_{ab}(s)}{\det{\bar{G}_{ab}}(s)}+i Q^{\mu \nu} \frac{G^L_{ab}}{\det{G^L_{ab}}} \ ,
\end{equation}
with
\begin{equation}\label{eqn7}
\bar{G}_{ab}(s)\equiv M_{ab}^2-\delta_{ab} s=
\left(\begin{array}{cc}
m_b^2+\Pi_b(s)-s & -\Pi_{ab}(s) \\
-\Pi_{ab}(s) & m_a^2+\Pi_a(s)-s
\end{array}\right) \ ,
\end{equation}
and
\begin{equation}\label{eqn7a}
G^L_{ab}(s)=\left(\begin{array}{cc}
\frac{m_b^2-s}{\Delta_b}+B_b(s) & -B_{ab}(s) \\
-B_{ab}(s) & \frac{m_a^2-s}{\Delta_a}+B_a(s)
\end{array}\right) \ ,
\end{equation}
where $M_{ab}^2$ is the mass matrix. After diagonalization, the mass
matrix becomes
\begin{equation}\label{eqn8}
M_{AB}^2=R M_{ab}^2 R^{\dagger}=
\left(\begin{array}{cc}
m_B^2 & 0 \\
0 & m_A^2
\end{array}
\right) \ .
\end{equation}
Note that the longitudinal term $G^L_{ab}/\det{G^L_{ab}}$ is nonvanishing,
but the poles are only related to the transverse term $\bar{G}_{ab}$.
By searching for the poles in the propagator matrix $G^{\mu
\nu}(s)$, which is equivalent to set $\det{\bar{G}(s)}=0$, we can
obtain the masses and widths of the physical states. In general,
there are two solutions $s_A$ and $s_B$ for the two state system. We
can also extract the mixing angle $\theta_{A,B}$ and the relative
phase angle $\phi_{A,B}$. These mixing parameters are different for
these two states, since they are extracted at the physical masses of
these two states, respectively. If $G$ is a normal matrix, which
means $G G^{\dagger}=G^{\dagger}G$, then it can be diagonalized
through a unitary transformation $R$. The resulting mixing angle
$\theta$ and relative phase $\phi$ can thus be uniquely determined.
Otherwise, we can only get a quasi-diagonalized matrix through the
unitary transformation $R$. The reason is because that orthogonality
cannot be satisfied between these two physical states.
\section{Coupling form factors in the chiral quark model}\label{sec:vertices}
At hadronic level all the vertices in Fig.~\ref{mixterm} involve the
Axial-Vector-Pseudoscalar (AVP) type of coupling. In general,
the AVP coupling vertex contains two coupling constants $g_S$ and
$g_D$ representing the $S$ and $D$ waves as shown in
Fig.~\ref{vertex}.
\begin{figure}[htbp]
\includegraphics[scale=0.7]{fig3}\\
\vspace{-17cm}
\caption{The AVP vertex via the $S$ and $D$ wave couplings.}\label{vertex}
\end{figure}
Since the decay momentum is small near the threshold, we expect that
contributions from the $D$-wave coupling would be small. As a
reasonable approximation, we omit $g_D$ and keep $g_S$ to the order
$O(v^0)$. In the multipole approach, the helicity amplitude for $1^+
\rightarrow 1^- + 0^-$ takes the form \cite{Zou:2002yy}
\begin{equation}\label{eqn9}
A_{\nu}=\langle S_f,\nu;0,0| \hat{S} | S_i,\nu \rangle = \sum_L \langle L, 0; S_f,\mu | S_i, \nu \rangle Y_{L0}(\hat{q})
G_L \ ,
\end{equation}
where $G_L$ is the coupling constant for the $L$ wave and $\hat{q}$
is the momentum direction of the final state particle in the center
of mass frame of the initial state. In the present case,
Eq.~(\ref{eqn9}) becomes
\begin{equation}\label{eqn10}
\left[
\begin{array}{c}
A_0 \\
A_1
\end{array}
\right]
=
\left[\begin{array}{cc}
\frac{1}{2 \sqrt{\pi}} & -\frac{1}{\sqrt{2 \pi}} \\
\frac{1}{2 \sqrt{\pi}} & \frac{1}{2 \sqrt{2 \pi}}
\end{array}
\right]
\left[\begin{array}{c}
G_S \\
G_D
\end{array}
\right] \ .
\end{equation}
In order to obtain $g_S$ to the order $O(v^0)$, we set
\begin{equation}\label{eqn11}
g_S=A_0(\vec{q} \to 0) =A_1(\vec{q} \to 0) \ .
\end{equation}
\begin{figure}[htbp]
\includegraphics[scale=0.7]{fig4}\\
\vspace{-16cm}
\caption{Pseudoscalar (a) and vector meson (b) emission via an active light quark $j$ in an effective chiral quark model.}\label{pic4}
\end{figure}
\subsection{Coupling to $D^* K$ and $D_{s}^{*} \eta$}
One notices that at all the coupling vertices the interacting quarks
involve only light quark, i.e. $u$, $d$ and $s$. By treating the
light mesons, pseudoscalar and vector mesons, as induced fields by a
chiral Lagrangian for the mesons coupling to constituent
quarks~\cite{Manohar:1983md}, the light and heavy quark degrees of
freedom can be separated out in terms of nonrelativistic expansions
near the decay threshold. This approach has been successfully
applied to light meson productions in photo-nucleon and
meson-nucleon
scatterings~\cite{Li:1997gd,Zhao:1998rt,Zhao:1998fn,Zhao:2002id,Zhao:2000iz,Zhong:2007fx,Zhong:2011ti}
and strong decays of heavy-light
mesons~\cite{Zhong:2008kd,Zhong:2009sk} recently.
In the chiral quark model, we treat the pseudoscalar mesons $K$ and
$\eta$ as the effective chiral fields as shown in
Fig.~\ref{pic4}(a). For emitting a pseudoscalar from an active quark
line, the quark-meson coupling and corresponding non-relativistic
form are respectively as follows~\cite{Zhong:2008kd}:
\begin{eqnarray}
\label{eqn12.1}
H_m &=& \sum_j \frac{1}{f_m} \hat{I}_j \bar{\psi}_j \gamma^{j}_{\mu} \gamma^{j}_{5} \psi_{j} \partial^{\mu} \phi_m \ ,\\
\label{eqn12.2}
H^{nr}_{m} &=& \sum_j \frac{1}{f_m}
\left\{
G \boldsymbol{\sigma}_j \cdot \textbf{q} + h \boldsymbol{\sigma}_j \cdot \textbf{p}^{i'}_j
\right\}
\hat{I}_j \exp(-i \textbf{q} \cdot \textbf{r}_j) \ ,
\end{eqnarray}
with
\begin{equation}\label{eqn13}
G\equiv -\left(1+\frac{\omega}{E_f+M_f}\right),
\quad
h\equiv \frac{\omega}{2 \mu_q} \ ,
\end{equation}
where $f_m$ is the decay constant of the pseudoscalar meson,
$\hat{I}_j$ the isospin operator, $\omega$ the energy of the
pseudoscalar, $M_f$ and $E_f$ the mass and energy of the final state
heavy meson, $\mu_{q}$ a reduced mass given by $1/\mu_{q}\equiv
1/m_j+1/m_j'$, $\textbf{p}^{i'}_j$ and $\textbf{r}_j$ the internal
momentum and coordinate for the light ($j$th) quark of the final
state heavy meson.
Following the procedure in \cite{Zhong:2008kd}, we derive the
helicity amplitude $A^{q}_{\nu}\equiv \langle S_f, \nu| \hat{H}_m
|S_i, \nu \rangle$ in the quark level. For $1{}^3P_1 \to 1{}^3S_1 +
\mathbb{P}$, the explicit expressions are
\begin{equation}\label{eqn14}
A^q_0=i g_1 h \alpha \exp(-\frac{q_1^2}{4 \alpha^2}),\quad
A^q_1=i \frac{g_1}{4 \alpha} \left[2 G q q_1 + h (4 \alpha^2-q_1^2)\right] \exp(-\frac{q_1^2}{4
\alpha^2}) \ ,
\end{equation}
and for $1{}^1P_1 \to 1{}^3S_1 + \mathbb{P}$, we have
\begin{equation}\label{eqn15}
A^q_0=-i \frac{g_1}{2\sqrt{2} \alpha}\left[2G q q_1 + h (2\alpha^2-q_1^2) \right] \exp(-\frac{q_1^2}{4 \alpha^2}),\quad
A^q_1=- \frac{i}{\sqrt{2}}g_1 h \alpha \exp(-\frac{q_1^2}{4
\alpha^2}) \ ,
\end{equation}
where $g_1=\langle \mathbb{M}_f | \hat{I}_1| \mathbb{M}_i \rangle$
is the isospin factor, $\alpha$ the harmonic oscillator strength
$\alpha\equiv \beta \left( {2 m_2}/(m_1+m_2) \right)^{{1}/{4}}$ as
in Ref.~\cite{Zhong:2008kd}, and $q_1\equiv q m_2/(m_1+m_2)$. In the
$c\bar{s}$ system, the 1st quark is $\bar{s}$ and the 2nd is $c$
quark, and the flavor symmetry between the heavy and light quark is
apparently broken.
By taking equivalence between the quark and hadron level helicity
amplitudes, we can extract the coupling form factor as follows:
\begin{equation}\label{eqn16}
A_{\nu}=\sqrt{(E_i+M_i)(E_f+M_f)} A^{q}_{\nu} \ .
\end{equation}
Then from Eqs.(\ref{eqn11}), (\ref{eqn14}) and (\ref{eqn15}), we
finally obtain:
\begin{eqnarray}
\label{eqn17.1}
\textrm{for}\quad 1{}^3P_1, & & g_S=-\frac{\delta}{f_m} \sqrt{2M_i(E_f+M_f)} \cdot g_1 h \alpha \exp(-\frac{q_1^2}{4 \alpha^2}) \ ,\\
\label{eqn17.2}
\textrm{for}\quad 1{}^1P_1, & & g_S= \frac{\delta}{f_m} \sqrt{2M_i(E_f+M_f)} \cdot \frac{1}{\sqrt{2}} g_1 h \alpha \exp(-\frac{q_1^2}{4
\alpha^2}) \ ,
\end{eqnarray}
where $\delta$ is a global parameter accounts for the strength of
the quark-meson couplings as introduced in \cite{Zhong:2008kd}.
\subsection{Coupling to $D K^*$}
In this coupling, the vector meson $K^*$ is treated as an effective
chiral field, for which the effective quark-vector-meson coupling
Lagrangian and the corresponding non-relativistic coupling
form~\cite{Zhao:1998fn,Riska:2000gd} are
\begin{eqnarray}
\label{eqn18.1}
\hat{H}_v &=& \sum_j a \bar{\psi}_j \gamma_{\mu}^j \phi_v^{\mu} \psi_j \ ,\\
\label{eqn18.2}
\hat{H}_v^T &=& \sum_j \left\{ -\frac{\textbf{p}_j^{i'} \cdot \boldsymbol{\epsilon}^*}{2 \mu_q} + i \boldsymbol{\sigma}_j \cdot \textbf{q} \times \boldsymbol{\epsilon}^* \left( \frac{1}{2 m_j}+\frac{1}{E_f+M_f}-\frac{m_j'}{2 M' m_j} \right)
\right\} a \hat{I}_j \exp(-i \textbf{q} \cdot \textbf{r}_j) \ ,\\
\label{eqn18.3}
\hat{H}_v^L &=& \sum_j \left\{ \left[ \frac{q}{\mu}(1-\frac{\omega}{2 m_j})+ \frac{q \omega}{2 M' \mu} + \frac{q \omega m_j'}{2 M' \mu m_j} \right]-\frac{\omega}{2 \mu \mu_q} \textbf{p}_j^{i'} \cdot \hat{\textbf{q}}
\right\} a \hat{I}_j \exp(-i \textbf{q} \cdot \textbf{r}_j) \ ,
\end{eqnarray}
where $\mu$, $\omega$ and \textbf{$\epsilon$} are the mass, energy
and polarization vector of the emitted vector meson, $M'$ the sum
of the constituent quark mass of the final meson, $a$ the overall
quark-vector-meson coupling, and other symbols have the same meaning
as those in Eqs.(\ref{eqn12.1})-(\ref{eqn13}). Using the above
operators, we can extract the helicity amplitudes $A_{\nu}^q$, i.e.
for $1{}^3P_1 \to 1{}^1S_0 + \mathbb{V}$,
\begin{equation}\label{eqn19}
A_1^q=-\frac{i}{2 \sqrt{2}} a g_1 B \frac{q_1}{\alpha}
\exp(-\frac{q_1^2}{4 \alpha^2}) ,\quad A_0^q= 0 \ ,
\end{equation}
and for $1{}^1P_1 \to 1{}^1S_0 + \mathbb{V}$,
\begin{equation}\label{eqn20}
A_1^q= i a g_1 A \alpha \exp(-\frac{q_1^2}{4 \alpha^2}) ,\quad
A_0^q= \frac{i}{\sqrt{2}} a g_1 \left[ - C \frac{q_1}{\alpha} + D
\alpha \left(\frac{q_1^2}{2 \alpha^2} -1\right) \right]
\end{equation}
with
\begin{eqnarray}
\label{eqn21.1}
A &\equiv & -\frac{1}{2\sqrt{2} \mu_q} \ ,\\
\label{eqn21.2}
B &\equiv & -\sqrt{2} q \left( \frac{1}{2 m_j} + \frac{1}{E_f + M_f} - \frac{m_j'}{2M' m_j} \right) \ ,\\
\label{eqn21.3}
C &\equiv & -\left[ \frac{q}{\mu}(1-\frac{\omega}{2 m_j})+ \frac{q \omega}{2 M' \mu} + \frac{q \omega m_j'}{2 M' \mu m_j} \right] \ ,\\
\label{eqn21.4}
D &\equiv & \frac{\omega}{2 \mu \mu_q} \ .
\end{eqnarray}
Substituting Eqs.~(\ref{eqn19}) and (\ref{eqn20}) to
Eqs.~(\ref{eqn11}) and (\ref{eqn16}),
we obtain
\begin{eqnarray}
\label{eqn22.1}
\textrm{for} \quad 1{}^3P_1, && g_S=0 \ ,\\
\label{eqn22.2}
\textrm{for} \quad 1{}^1P_1, && g_S=- \sqrt{(E_i+M_i)(E_f+m_f)} \cdot \frac{1}{\sqrt{2}} a g_1 \frac{\alpha}{\mu_q} \exp(-\frac{q_1^2}{4
\alpha^2}) \ .
\end{eqnarray}
The coupling $g_S=0$ in Eq.~(\ref{eqn22.1}) is because $A_1^q$ in
Eqs.~(\ref{eqn19}) is proportional to $q_1$. Thus, the effective
coupling vanishes below the open decay threshold. As a consequence,
the contributions from Fig.~\ref{mixterm}(4) and (5) should vanish
to the order $O(v^0)$. Therefore, we only need to consider the
contributions from Fig.~\ref{mixterm}(1-3) in the following
calculations.
\subsection{Numerical results for couplings}
In the numerical calculation, we set $|\vec{q}|=0$ when the initial
state lies below the threshold for $\mathbb{V P}$ final state
\cite{Riska:2000gd}. We adopt
$\eta=\frac{1}{\sqrt{3}}(u\bar{u}+d\bar{d}-s\bar{s})$ in the
$\eta-\eta'$ mixing scheme which corresponds to
$\theta_P=-\arcsin(1/3)=-19.47^\circ$ for the flavor octet and
singlet mixing. Since the contribution from the $D_s^*\eta$ loop is
small, the uncertainties with $\theta_P$ have only negligible
effects on the mixing matrix. We obtain isospin factors $g_1$ for
different intermediate states as listed in Table~\ref{tab2}.
\begin{table}[htbp]
\centering
\caption{Isospin factors $g_1$ extracted in the quark model.}\label{tab2}
\begin{tabular}{c|c|c|c}
\hline\hline
$1{}^3P_1 / 1{}^1P_1$ & $D^{*0} K^+$ & $D^{*+} K^0$ & $D_s^{*} \eta$ \\
\hline
$g_1$ & 1 & 1 & $-\frac{1}{\sqrt{3}}$ \\
\hline\hline
\end{tabular}
\end{table}
The following values are adopted for other
parameters~\cite{Zhong:2008kd}: $\delta=0.557$, $\beta=0.4 \
\textrm{GeV}$, $f_K=f_{\eta}=160 \ \textrm{MeV}$, and the
constituent quark masses $m_u=m_d=350 \ \textrm{MeV}$, $m_s=550 \
\textrm{MeV}$, $m_c=1700 \ \textrm{MeV}$. We note that our numerical
results are not sensitive to $m_c=1500\sim 1700$ MeV, while the
light quark masses $m_u=m_d=330\sim 350$ MeV and $m_s=500\sim 550$
MeV will lead to about $(1\sim 5)\%$ uncertainties with the final
results.
The masses $M_i$ of the initial states $1{}^3 P_1$ and $1{}^1 P_1$
$c \bar{s}$ still have uncertainties. Fortunately, the couplings
$|g_S|$ change only $5\%$ at most when $M_i \in [2.460, \ 2.536] \
\textrm{GeV}$ as shown in Fig.~\ref{g_s}. Also it shows that the
couplings to $D^{*0} K^+$ and $D^{*+} K^0$ are almost the same for
each state due to the isospin symmetry. A set of typical $g_s$
couplings is listed in Table~\ref{tab3}.
\begin{figure}[htbp]
\includegraphics[scale=1]{fig5}\\
\vspace{0cm}
\caption{the absolute values of couplings $g_S$ as functions of the initial meson mass $M_i$}\label{g_s}
\end{figure}
\begin{table}[htbp]
\centering
\caption{Vertex couplings $g_S$ at $M_i=2.5$ GeV.}\label{tab3}
\begin{tabular}{c|c|c|c}
\hline\hline
$g_S$ (GeV) & $D^{*0} K^+$ & $D^{*+} K^0$ & $D_s^{*} \eta$ \\ \hline
$1{}^3P_1$ & $-7.982 $ & $-8.052 $ & $2.040 $ \\ \hline
$1{}^1P_1$ & $5.644 $ & $5.694 $ & $-1.443 $ \\
\hline\hline
\end{tabular}
\end{table}
Apart from the on-shell coupling $g_S$, the chiral quark model also
provides an exponential momentum-dependent form factor
$\exp(-{q_1^2}/4 \alpha^2)$ as shown in Eqs.~(\ref{eqn17.1}),
(\ref{eqn17.2}), (\ref{eqn22.1}), and (\ref{eqn22.2}). In order to
keep this feature in the meson loops, we modify the exponential form
factor to a covariant form:
\begin{equation}\label{eqn23}
\exp(-\frac{q_1^2}{4 \alpha^2}) \rightarrow
\exp(\frac{q^2-m^2}{\Lambda^2}) \ ,
\end{equation}
where $q^{\mu}$ and $m$ are the four-vector momentum and mass of
either $\mathbb{V}$ or $\mathbb{P}$ particle. Parameter $\Lambda$ is
the cut-off energy, which can be determined by the quark model,
namely, for the $D^* K$ and $D_s^* \eta$ loops,
\begin{eqnarray}\label{eqn24}
\Lambda=2 \frac{m_c+m_s}{m_c} \left[\frac{2 m_c}{m_c+m_s}\right]^{\frac{1}{4}} \beta = 1.174 \
\textrm{GeV} \ .
\end{eqnarray}
The exponential form factor serves to remove the ultraviolet
divergence in the loop integrals.
\section{The propagator matrix}\label{sec:propagator}
In this Section we will determine the propagator matrix $G$. From
Eqs.~(\ref{eqn17.1}) and (\ref{eqn17.2}), we have
\begin{equation}\label{eqn25}
\Pi_a=-\sqrt{2}\Pi_{ab}, \quad \Pi_b=-\Pi_{ab}/ \sqrt{2}.
\end{equation}
So, we only need to calculate the mixing term $\Pi_{ab}$. With the
AVP coupling form factors, we can explicitly write down $D_{ab}^{\mu
\nu}$ as the following:
\begin{eqnarray}\label{eqn26}
D_{ab}^{\mu \nu} &=& \sum g_a g_b \cdot i \int \frac{\ud^4 k}{(2\pi)^4} \frac{ \exp\left(\frac{k^2-m_v^2}{\Lambda^2}\right) \exp\left(\frac{(k+p)^2-m_p^2}{\Lambda^2}\right) }{[k^2-m_v^2][(k+p)^2-m_p^2]} \left(g^{\mu \nu}-\frac{k^{\mu}k^{\nu}}{m_v^2}\right) \nonumber \\
&\equiv & \sum g_a g_b(\Pi g^{\mu \nu} + B' p^{\mu}p^{\nu}/p^2) \equiv \sum g_a g_b
loop \ ,
\end{eqnarray}
where $g_a$ and $g_b$ are the $S$-wave couplings of the two
vertices, respectively. Comparing with Eq.~(\ref{eqn5}), we obtain
\begin{equation}\label{eqn27}
\Pi_{ab}\equiv \sum g_a g_b \Pi \ .
\end{equation}
The mixing term $\Pi_{ab}$ can be decomposed into two terms, i.e.
\begin{equation}\label{eqn31}
\Pi_{ab}\equiv\Pi_{ab}^1+\Pi_{ab}^2,
\end{equation}
with
\begin{equation}
\Pi_{ab}^1\equiv \sum g_a g_b \Pi^1, \quad \Pi_{ab}^2\equiv \sum g_a
g_b \Pi^2 \ ,
\end{equation}
where $\Pi_{ab}^1$ and $\Pi_{ab}^2$ are contributions from the
$g^{\mu \nu}$ and $k^{\mu}k^{\nu}$ terms of the vector propagator,
respectively.
As follows, we first make an on-shell approximation to investigate
the absorptive part. Then, we investigate the full integrals with
the help of the exponential form factors.
\subsection{On-shell approximation}
Since the absorptive part of a two-point function is independent of
the form factors, the on-shell approximation will allow us to
separate out the absorptive part and then compare it with that in a
full loop integral. Here we only consider $\Pi^1$,
for which the loop integral of Eq.~(\ref{eqn26}) in
the on-shell approximation becomes
\begin{equation}\label{eqn28}
loop_1 \stackrel{\textrm{on shell}}{\rightarrow} g^{\mu \nu} \frac{-i}{16 \pi^2} \textrm{Im} B0(s,m_p^2,m_v^2) =\Pi^1 g^{\mu
\nu} \ .
\end{equation}
The resulting mixing term $\Pi_{ab}$ is a function of $s$. We plot
$\Pi_{ab}(\sqrt{s})$ in Fig.~\ref{Dab} with the couplings listed in
Table~\ref{tab3} adopted.
\begin{figure}[htbp]
\includegraphics[scale=1]{fig6}\\
\vspace{0cm}
\caption{The mixing term $\Pi_{ab}$ in the on-shell approximation. }\label{Dab}
\end{figure}
In Fig.~\ref{Dab}, two kink structures can be identified. The first
one at $\sqrt{s}=2.501 \ \textrm{GeV}$ corresponds to the $D^{*0}
K^+$ threshold, and the second one at $\sqrt{s}=2.508 \
\textrm{GeV}$ to the $D^{*+} K^0$ threshold. This result will be
compared with the absorptive part in the full loop integrals later.
\subsection{Full loop calculation with the exponential form factor}
In this Subsection we perform the full loop calculation with the
exponential form factor. The explicit formula for $\Pi^1$ is
\begin{eqnarray}\label{eqn29}
loop_1&=&
i\int \frac{\ud^4 k}{(2\pi)^4} \frac{ \exp\left(\frac{k^2-m_v^2}{\Lambda^2}\right) \exp\left(\frac{(k+p)^2-m_p^2}{\Lambda^2}\right) }{[k^2-m_v^2][(k+p)^2-m_p^2]}
g^{\mu \nu} \nonumber \\
&=&
g^{\mu \nu} \frac{-1}{16 \pi^2} \int_0^1 \ud x \, e^c U(2,1,\frac{b^2}{a},a \Delta)
=\Pi^1 g^{\mu \nu}
\end{eqnarray}
with
\begin{eqnarray*}
a &\equiv & \frac{2}{\Lambda^2} \ ,\\
b^2 &\equiv & \frac{ (1-2x)^2 }{ \Lambda^4 } s \ , \\
c &\equiv & \frac{ s( 2x^2 - 2x + 1 )-m_p^2-m_v^2 }{ \Lambda^2 } \ , \\
\Delta &\equiv & (1-x)m_v^2 + x m_p^2 -x(1-x)s \ .
\end{eqnarray*}
The explicit formula for $\Pi^2$ is
\begin{eqnarray}\label{eqn32}
loop_2&=&
(-i)\int \frac{\ud^4 k}{(2\pi)^4} \frac{ \exp\left(\frac{k^2-m_v^2}{\Lambda^2}\right) \exp\left(\frac{(k+p)^2-m_p^2}{\Lambda^2}\right) }{[k^2-m_v^2][(k+p)^2-m_p^2]}
\frac{k^{\mu}k^{\nu}}{m_v^2} \nonumber \\
&=&
g^{\mu \nu} \frac{1}{32 \pi^2 a m_v^2} \int_0^1 \ud x \, e^c U(2,0,\frac{b^2}{a},a \Delta) + B_2^U p^{\mu}p^{\nu} \nonumber \\
&\equiv &
\Pi^2 g^{\mu \nu} + B_2 p^{\mu}p^{\nu} \ ,
\end{eqnarray}
where $a, \ b, \ c, \ \Delta$ are the same as those in
Eq.~(\ref{eqn29}). The function $U(a,b,c,z)$ is a class of special
integrals which appears in the evaluation of the loop integrals with
exponential form factors. The detailed definition and calculation of
$U(a,b,c,z)$ are provided in Appendix~\ref{app:1}.
The full loop calculation of the mixing term $\Pi_{ab}(\sqrt{s})$ is
presented in Fig.~\ref{DabExp} where the parameters are the same as
before. In order to see clearly the contributions from different
parts, we also give two sets of the calculated values in
Tables~\ref{tab:2460} and \ref{tab:2536}.
\begin{figure}[htbp]
\includegraphics[scale=1]{fig7}\\
\vspace{0cm}
\caption{(color online). The mixing term $\Pi_{ab}$ with exponential form factors.
$\Pi_{ab}^1$ and $\Pi_{ab}^2$ are the contributions from the $g^{\mu \nu}$ term and $k^{\mu}k^{\nu}$ term of the vector propagator, respectively.
The dashed lines represent the dispersive parts, while the dot-dashed lines represent the absorptive ones.}\label{DabExp}
\end{figure}
\begin{table}[htbp]
\centering
\caption{The mixing term $\Pi_{ab}$ at the pole position $\sqrt{s}=2.4545 \ \textrm{GeV}$. }\label{tab:2460}
\begin{tabular}{c||c|c|c||c}
\hline \hline
intermediate state & $D^{*0}K^+$ & $D^{*+} K^0$ & $D_s^* \eta$ & $\Pi_{ab}=\sum g_a g_b \Pi \ ({\textrm{GeV}}^2)$ \\ \hline
$g_a g_b ({\textrm{GeV}}^2)$ & $-45.05$ & $-53.90$ & $-2.944$& ---\\ \hline
$\Pi^1$(on-shell) & $0$ & $0$ & $0$& $0$\\ \hline
$\Pi^1$ & $-2.391\times 10^{-3}$ & $-2.253 \times 10^{-3}$ & $-0.868 \times 10^{-3}$ & $0.2317$ \\ \hline
$\Pi^2$ & $5.650 \times 10^{-5}$ & $5.448 \times 10^{-5}$ & $2.503 \times 10^{-5}$ & $-0.0056$ \\
\hline \hline
\end{tabular}
\end{table}
\begin{table}[htbp]
\centering
\caption{The mixing term $\Pi_{ab}$ at the pole position $\sqrt{s}=(2.5449-0.0010 i)\ \textrm{GeV}$.}\label{tab:2536}
\begin{tabular}{c||c|c|c||c}
\hline \hline
intermediate state & $D^{*0}K^+$ & $D^{*+} K^0$ & $D_s^* \eta$ & $\Pi_{ab}=\sum g_a g_b \Pi \ ({\textrm{GeV}}^2)$ \\ \hline
$g_a g_b ({\textrm{GeV}}^2)$ & $-45.05$ & $-53.90$ & $-2.944$& ---\\ \hline
$\Pi^1$(on-shell) & $-2.969i\times 10^{-3}$ & $-2.718 i\times 10^{-3}$ & $0$ & $0.2803i$ \\ \hline
$\Pi^1$ & $(-4.157-2.969i)\times 10^{-3}$ & $(-4.260-2.718i)\times 10^{-3}$ & $-1.389\times 10^{-3}$ & $0.4210+0.2803i$ \\ \hline
$\Pi^2$ & $(9.419+0.886i) \times 10^{-5}$ & $(9.059+0.678i )\times 10^{-5}$ & $3.560 \times 10^{-5}$ & $-0.0092-0.0008i$ \\ \hline
\hline
\end{tabular}
\end{table}
The loop calculation results help us to learn the following points:
\begin{itemize}
\item The imaginary part of $\Pi_{ab}^1$ with exponential form factors
is the same as that in the on-shell approximation. It justifies our
calculation method for $U(a,b,c,z)$ as described in the Appendix.
\item The contribution from the term of $g^{\mu \nu}$ is dominant. The open thresholds
of $D^{*0} K^+$ and $D^{*+}K^0$ cause two kinks in both real and
imaginary parts. With the increase of $\sqrt{s}$, $\textrm{Re}
\Pi_{ab}^1$ first increases until it reaches a summit at the $D^{*+}
K^0$ threshold. It then decreases in a linear behavior in terms of
$\sqrt{s}$. In contrast, $\textrm{Im} \Pi_{ab}^1$ is zero below the
$D^{*0} K^+$ threshold and then increases quickly when the decay
thresholds are open. One can see that below the $D^{*0} K^+$
threshold, the real part is the only contribution and cannot be
neglected. The imaginary part becomes significant above 2.53 GeV.
\item The calculation also shows that the contributions from the $k^{\mu} k^{\nu}$
term of the vector propagator are negligible. Near the threshold,
the momentum is small such that $\Pi^2_{ab}$ suffers an $O(1/m_v^2)$
suppression comparing to $\Pi^1_{ab}$ in both the absorptive and
dispersive part.
\item The contributions from the $D^* K$ loops are found dominant, while the
contributions from $D_s^* \eta$ account for only about $1\%$ of the
mixing term due to the rather small coupling value in the $D_s^*
\eta$ loop.
\end{itemize}
\section{Pole positions and mixing parameters }\label{sec:mass}
With the $\Pi_{ab}(s)$ determined, we can directly search for poles
for the physical states in the propagator matrix $G$ in
Eq.~(\ref{eqn6}). We adopt the following bare $c\bar{s}$ masses,
$m[{}^3P_1]=2.57 \ \textrm{GeV}$ and $m[{}^1P_1]=2.53 \
\textrm{GeV}$, from the Godfrey-Isgur (GI)
model~\cite{Godfrey:1985xj} as input. By scanning over the energy
$\sqrt{s}$, the requirement of $|\det[\bar{G}(s)]|=0$ provides a
direct access to the pole positions as shown in Fig.~\ref{poles}.
Two possible poles near 2.46 GeV and 2.54 GeV are highlighted. When
varying the cut-off parameter $\Lambda$ in Eq.~(\ref{eqn24}) within
the range of $[1.174-0.22,1.174+0.22]$ GeV , it shows that the
higher pole is stable and the lower one changes from 2.47 GeV to
2.44 GeV. Searching for the poles on the complex energy plane, we
can pin down the masses and widths of these two poles as listed in
Table~\ref{tab:mass}. It shows that the mass of $D_{s1}(2460)$
changes 3.6 MeV at most with or without the contribution from the
$k^{\mu}k^{\nu}$ term of the propagator, while the mass of
$D_{s1}(2536)$ changes only 0.1 MeV. The extracted mass of
$D_{s1}(2460)$ is only 5 MeV below the experiment value, and the
mass of $D_{s1}(2536)$ is only 10 MeV above the experiment one. In
principle, the Okubo-Zweig-Iizuka (OZI) rule allowed hadronic decay
width of $D_{s1}(2460)$ is zero. The obtained width 2.0 MeV for
$D_{s1}(2536)$ seems to be slightly larger than the experiment value
0.92 MeV, but can still be regarded as in good agreement. In brief,
our prediction for the masses and widths of these two states agrees
well with the experiment data.
\begin{figure}[htbp]
\includegraphics[scale=1]{fig8}\\
\vspace{0cm}
\caption{Pole structures highlighted by the zero values of $\det[\bar{G}]$ in the propagator matrix. }\label{poles}.
\end{figure}
\begin{table}[htbp]
\centering
\caption{Masses and widths obtained from the pole analysis.}\label{tab:mass}
\begin{tabular}{c|c|c}
\hline \hline
$[m-i\frac{\Gamma}{2}]$ (MeV) & $D_{s1}(2460)$ & $D_{s1}(2536)$ \\ \hline
$\Pi_{ab}^1 $ & $2454.5$ & $2544.9-1.0 i$ \\ \hline
$\Pi_{ab}^1+\Pi_{ab}^2$ & $2455.8$ & $2544.9-1.1 i$ \\ \hline
Experiment & $2459.5$ & $2535.08-0.46 i$ \\
\hline \hline
\end{tabular}
\end{table}
Before extracting the mixing parameters, we show that our formalisms
can reproduce the ideal mixing angle $\theta_0$ in the heavy quark
limit. In this limit, $m_a$ and $m_b$ are degenerate. From
Eqs.~(\ref{eqn7}) and (\ref{eqn25}), we only need to diagonalize the
simple matrix
\begin{equation}\label{eqn33}
\left(
\begin{array}{cc}
-\frac{1}{\sqrt{2}} & -1 \\
-1 & -\sqrt{2} \\
\end{array}
\right) \ ,
\end{equation}
which immediately leads to
$\theta_0=\arctan[1/\sqrt{2}]=35.26^\circ$.
Now we proceed to the extraction of the mixing parameters
$\{\theta,\phi\}$ by diagonalizing $\bar{G}(s)$ with $\sqrt{s}$
fixed at the poles. When $\bar{G}$ is a complex matrix, we try to
approach the diagonal limit $R \bar{G}_{ab}R^{\dagger}=\bar{G}_{AB}$
in three ways: Method \Rmnum{1}, set $\bar{G}_{12}^{AB}=0$; Method
\Rmnum{2}, set $\bar{G}_{21}^{AB}=0$; and Method \Rmnum{3}, minimize
$|\bar{G}_{12}^{AB}| + |\bar{G}_{21}^{AB}|$. The results from these
three diagonalization schemes are listed in Table~\ref{tab:mixing}.
As we expected before, the mixing angles of these two states
determined at their pole masses are indeed different. For
$D_{s1}(2460)$, $\bar{G}$ is a symmetric real matrix. So the mixing
parameters are the same in these three methods: $\theta=47.6^\circ,
\, \phi=0^\circ$. From the mixing scheme in Eq.~(\ref{eqn2:1}),
$\theta>45^\circ$ means that the ${}^1P_1$ component is larger than
the ${}^3P_1$ in $D_{s1}(2460)$. This mixing pattern would affect
the mass shift as we will show later. The result corresponds to
$\theta'=12.3^\circ$ in the $j=1/2$ and $j=3/2$ mixing in the heavy
quark limit. For $D_{s1}(2536)$ , $\bar{G}$ is a complex matrix. The
mixing angle $\theta=39.7^\circ$ determined at the $D_{s1}(2536)$
mass changes little in those three methods, while the relative phase
suffers an uncertainty of $\phi=-6.9^\circ \sim 6.9^\circ$. We will
show later in Sec.~\ref{sec:exp} that the mixing angle
$\theta=39.7^\circ$ is consistent with the experimental constraints
and can be useful for picking up one of those two solutions from the
experimental fit. Again from the mixing scheme, $\theta<45^\circ$
means that the ${}^1P_1$ component is larger than the ${}^3P_1$ one
in $D_{s1}(2536)$. The result corresponds to $\theta'=4.4^\circ$ in
the $j=1/2$ and $j=3/2$ mixing bases. The energy dependence of the
mixing angle reflects the breaking of orthogonality among these two
physical states.
\begin{table}[htbp]
\centering
\caption{The mixing angle $\theta$ and relative phase $\phi$ extracted at the two poles in those three diagonalization schemes.}\label{tab:mixing}
\begin{tabular}{|c||c|c|c||c|c|c|}
\hline
& \multicolumn{3}{c||}{$D_{s1}(2460)$} & \multicolumn{3}{c|}{$D_{s1}(2536)$} \\ \cline{2-7}
\raisebox{2.3ex}[0pt]{$\{\theta,\phi\}[{}^{\circ}]$}& \Rmnum{1} & \Rmnum{2} & \Rmnum{3} & \Rmnum{1} & \Rmnum{2} & \Rmnum{3} \\ \hline
$\Pi_{ab}^1 $ & $\{47.5, \ 0\}$ & $\{47.5, \ 0\}$ & $\{47.5, \ 0\}$ & $\{39.7, \ -6.4\}$ & $\{39.7, \ 6.4\}$ & $\{39.7, \ 0\}$ \\ \hline
$\Pi_{ab}^1+\Pi_{ab}^2$ & $\{47.6, \ 0\}$ & $\{47.6, \ 0\}$ & $\{47.6, \ 0\}$ & $\{39.8, \ -6.5\}$ & $\{39.8, \ 6.5\}$ & $\{39.7, \ 0\}$ \\
\hline
\end{tabular}
\end{table}
From the mixing angle analysis, we also learn that the
$D_{s1}(2460)$ has a larger $j=1/2$ component which couples to the
$D^* K$ through an $S$-wave. It hence acquires a significant mass
shift $\sim 100$ MeV through meson loop corrections. In contrast,
the $D_{s1}(2536)$ contains a larger $j=3/2$ component which couples
to the $D^*K$ through a $D$-wave. It only gains a small mass shift
$\sim 10$ MeV.
\begin{figure}[hbtp]
\includegraphics[scale=1]{fig9}
\vspace{0cm}
\caption{(color online). Schematic plot for the mass-shift
procedure. The thin solid bars represent the original ${}^3P_1$ and
${}^1P_1$ states in the quark model. The thick solid bars represent
the two physical states $D_{s1}(2460)$ (left) and $D_{s1}(2536)$
(right). The solid arrows represent the mass shifts due to the
diagonal elements $\Pi_a$ and $\Pi_b$, while the dashed arrows
represent those due to the off-diagonal element $\Pi_{ab}$. The
threshold for $D^* K$ is shown by the horizontal dashed
line.}\label{massShift}
\end{figure}
\begin{table}[htbp]
\centering
\caption{Mass shift procedure at different $\sqrt{s}$. From $\sqrt{M_{ab}^2}$, we can see the mass shifts
due to $\Pi_a$ and $\Pi_b$, while from $\sqrt{M_{AB}^2}$, further mass shifts due to $\Pi_{ab}$ can be learned.
Here we only present the results from the quasi-diagonalization method \Rmnum{2}. }\label{tab:shift}
\begin{tabular}{cccc}
\hline \hline
$\sqrt{s}$ GeV & bare mass (GeV) & $\sqrt{M_{ab}^2}$ (GeV) & $\sqrt{M_{AB}^2}$ (GeV) \\ \hline
2.4545 GeV &
& $\left[\begin{array}{cc}
2.497 & 0.481 i \\
0.481 i & 2.505
\end{array}\right]$
& $\left[\begin{array}{cc}
2.548 & 0 \\
0 & 2.455
\end{array}
\right]$
\\
2.5449 GeV &
\raisebox{3ex}[-10pt]
{$\left[\begin{array}{cc}
2.53 & 0 \\
0 & 2.57
\end{array}\right]$}
& $\left[\begin{array}{cc}
2.471-0.040i & 0.206-0.678i \\
0.206-0.678i & 2.454-0.080i
\end{array}
\right]$
& $\left[\begin{array}{cc}
2.545-0.001i & 0.141+0.303i \\
0 & 2.379-0.123i
\end{array}
\right]$ \\
\hline \hline
\end{tabular}
\end{table}
The mass shift procedure is also an interesting issue and can help
us to understand why $D_{s1}(2460)$ has a larger ${}^1P_1$
component. As shown in Fig.~\ref{massShift} and
Table~\ref{tab:shift}, we can decompose the mass shift procedure
into two classes, i.e. diagonal shift and off-diagonal shift. The
diagonal elements $\Pi_a$ and $\Pi_b$ cause both ${}^3P_1$ and
${}^1P_1$ states to move downwards, while the off-diagonal elements
$\Pi_{ab}$ make one state to shift up and the other to shift down.
At $\sqrt{s}=2.46 \ \textrm{GeV}$, after the diagonal shift the
${}^3P_1$ state is still higher than the ${}^1P_1$. But after the
off-diagonal shift, the higher mass state moves down to become an
on-shell $D_{s1}(2460)$ and the lower state moves up to become a
virtual $D_{s1}(2536)$. The reversal of the mass ordering results in
a mixing angle $\theta > 45^\circ$ and thus a larger ${}^1P_1$
component in $D_{s1}(2460)$. At $\sqrt{s}=2.54 \ \textrm{GeV}$,
after the diagonal shift the ${}^1P_1$ becomes higher than the
${}^3P_1$. Then after the off-diagonal shift, the higher state
becomes much higher and the lower much lower, which causes a mixing
angle $\theta < 45^\circ$ and a larger ${}^1P_1$ component in
$D_{s1}(2536)$. Note that in this situation the on-shell state
corresponds to the $D_{s1}(2536)$, and the $D_{s1}(2460)$ appear as
a virtual one.
\section{Experimental constraints on the mixing angle}\label{sec:exp}
In this part, we come to survey the constraints for the mixing
angle $\theta$ from experiments. The strong decays of $D_{s1}(2536)$
has been measured with reasonable precision which are summarized in
Table~\ref{tab:expdata}. Since the $D^* K$ channel is the only
allowed strong decay channel for $D_{s1}(2536)$, it is a good
approximation to assume
\begin{equation}\label{eqn:exp1}
\Gamma[D_{s1}(2536)] \approx \Gamma( D_{s1}(2536)\to D^* K ) \ ,
\end{equation}
which can be estimated in the chiral quark model. The partial width
fractions $R_1$ and $R_2$ can also be calculated and compared with
the data.
\begin{table}[htbp]
\centering
\caption{ The available experimental status of $D_{s1}(2536)$. }\label{tab:expdata}
\begin{tabular}{c}
\hline\hline
$
m = 2535.08\pm 0.01 \pm 0.15 \ \textrm{MeV}, \quad
\Gamma = 0.92 \pm 0.03 \pm 0.04 \ \textrm{MeV} \quad (\textrm{BaBar \cite{Lees:2011um}})
$
\\ \hline
$
R_1 = \frac{ \Gamma (D^*(2007)^0 K^+ ) }{ \Gamma (D^*(2010)^+ K^0 ) } = 1.36 \pm 0.20 \quad (\textrm{PDG2010 \cite{Nakamura:2010zzi}})
$
\\ \hline
$
R_2 = \frac{ \Gamma (D^*(2010)^+ K^0 )_{S-wave} }{ \Gamma (D^*(2010)^+ K^0 ) } = 0.72 \pm 0.05 \pm 0.01 \quad(\textrm{Belle \cite{:2007dya}})
$
\\
\hline\hline
\end{tabular}
\end{table}
The helicity amplitudes for $ 1{}^3P_1 \to D^* K$ and $1{}^1P_1 \to
D^* K$ have been listed in Eqs.~(\ref{eqn14}) and (\ref{eqn15}). The
partial width can be obtained by~\cite{Zhong:2008kd}
\begin{equation}\label{eqn:exp2}
\Gamma =\left( \frac{\delta}{f_m} \right)^2 \frac{( E_f + M_f ) |\vec{q}| }{ 4 \pi M_i ( 2J_i + 1 ) } \sum_{\nu} \left| A_{\nu}^q
\right|^2 \ ,
\end{equation}
where $J_i$ is the spin of the initial particle. In order to
calculate $R_2$, we need to extract the $S$-wave components from the
helicity amplitudes. By defining $A_s= \frac{G_S}{ 2\sqrt{\pi} }$
and $A_D= \frac{G_D}{ 2\sqrt{\pi} }$, we deduce from
Eq.~(\ref{eqn10})
\begin{equation}\label{eqn:exp3}
\left\{\begin{array}{l}
A_0 = A_S - \sqrt{2} A_D \\
A_1 = A_S + \frac{1}{\sqrt{2}} A_D
\end{array}
\right.
\Rightarrow
\left\{\begin{array}{l}
A_S = \frac{1}{3}( A_0 + 2 A_1 ) \\
A_D = -\frac{\sqrt{2}}{3}( A_0 - A_1 )
\end{array}
\right. \ ,
\end{equation}
where the $S$ and $D$-wave components have been separated out. We
use the same model parameters as before to calculate the partial
width $\Gamma[D_{s1}(2536)]$ and ratios $R_1$ and $R_2$ in terms of
the mixing angle $\theta$. The results are shown in
Figs.~(\ref{fig:width})-(\ref{fig:exp_theta}).
\begin{figure}[htbp]
\begin{minipage}[b]{0.33\textwidth}
\includegraphics[width=0.9\textwidth]{fig10}\\
\caption{$\Gamma[D_{s1}(2536)]$ as a function of $\theta$}\label{fig:width}
\end{minipage
\begin{minipage}[b]{0.33\textwidth}
\includegraphics[width=0.9\textwidth]{fig11}\\
\caption{$R_1$ as a function of $\theta$}\label{fig:R_1}
\end{minipage
\begin{minipage}[b]{0.33\textwidth}
\includegraphics[width=0.9\textwidth]{fig12}\\
\caption{$R_2$ as a function of $\theta$}\label{fig:R_2}
\end{minipage}
\end{figure}
A similar result as Fig.~\ref{fig:width} for $\Gamma[D_{s1}(2536)]$
in terms of $\theta$ has been given in Ref.~\cite{Zhong:2008kd} but
with the notation $\theta \to \phi + 90^\circ$. Those three
horizontal lines in Figs.~(\ref{fig:width})-(\ref{fig:R_2})
represent the upper limits, center values, and lower limits of the
experimental data. The interesting feature arising from the results
of Figs.~(\ref{fig:width})-(\ref{fig:R_2}) is that the overlaps
between the experimental data and theoretical values are separated
into two narrow bands of $\theta$ which are located symmetric to the
ideal mixing angle $\theta_0=35.26^\circ$, i.e. $\theta_1\simeq
32.1^\circ$ or $\theta_2\simeq 38.4^\circ$. An alternative way to
present the results is via Fig.~\ref{fig:exp_theta}, where the
overlapped $\theta$ values are denoted by the vertical dashed lines,
while the experimental observables with errors are presented in
terms of $\theta$. Notice that these two bands of $\theta$ are both
smaller than $45^\circ$. Therefore, based on the present
experimental measurements, one cannot determine which value for
$\theta$ should be taken. It turns out that our analysis in
Sec.~\ref{sec:mass} can precisely pick up one of these two
solutions, namely, $\theta_2\simeq 38.4^\circ$ is favored in
comparison with the theoretical value $\theta=39.7^\circ$.
\begin{figure}[hbtp]
\includegraphics[scale=1]{fig13}
\vspace{0cm}
\caption{The experimental constraints on the mixing angle $\theta$.
}\label{fig:exp_theta}
\end{figure}
\section{summary}\label{sec:summary}
In summary, we have studied the mixing mechanism for the axial
vector states $D_{s1}(2460)$ and $D_{s1}(2536)$ via the $S$-wave
intermediate meson loops. We establish the propagator matrix for
this two-state system. Then, by searching for the pole structures in
the propagator matrix, we can pin down the masses and widths of the
physical states. The mixing angle and relative phase between the
${}^3P_1$ and ${}^1P_1$ components can be determined by
diagonalizing the propagator matrix. For $D_{s1}(2460)$, we obtain
$m=2454.5 \ \textrm{MeV},\, \theta=47.5^\circ$ and $\phi=0^\circ$.
For $D_{s1}(2536)$, we find $m=2544.9-1.0 i \ \textrm{MeV},\,
\theta=39.7^\circ$, and $\phi=-6.9^\circ \sim 6.9^\circ$. Our
results agree well with the experimental measurement. In particular,
the new BaBar measurement put a strong constraint on the mixing
angle at the mass of $D_{s1}(2536)$ with two solutions,
$\theta_1\simeq 32.1^\circ$ and $\theta_2\simeq 38.4^\circ$. Our
theoretical calculation finds $\theta=39.7^\circ$ which is in good
agreement with $\theta_2$.
Note that due to the breaking of orthogonality the energy-dependent
mixing angles defined at the different physical masses turn out to
have different values. We find that both $D_{s1}(2460)$ and
$D_{s1}(2536)$ have a relatively large ${}^1P_1$ component in their
wavefunctions.
It is also interesting to learn the important role played by the
coupled channel effects for states near open thresholds. For states
that can couple to each other via the coupled channels, the
two-state propagator matrix carries rich information about the
mixing and mass shifts as a manifestation of the underlying
dynamics. Extension of such a study to other axial-vector meson
mixings would be useful for deepen our understanding of the coupled
channel effects and their impact on the hadron spectrum.
\acknowledgments
This work is supported, in part, by National Natural Science
Foundation of China (Grant No. 11035006), Chinese Academy of
Sciences (KJCX2-EW-N01), and Ministry of Science and Technology of
China (2009CB825200).
|
\section{Introduction}
Recently there has been a flurry of interest in the galileon modification of gravity\cite{nicolis}\cite{fairlie1}\cite{fairlie2}. The theory is motivated by the DGP brane world model \cite{divali}. Galileon fields can be considered to be a generalization of the boundary effective field theory on the DGP brane at the so called \textit{Decoupling limit} \cite{Luty:2003vm} where it corresponds to the brane-bending mode (for a review see \cite{tonyrev}). The simplest version of galileon fields are postulated to have a novel symmetry structure, ie the action is invariant under $\pi \to \pi + a_\mu x^\mu + b$ ($a_\mu , b$ are constants). Demanding that the equation of motion only contains derivatives of order up to 2, gives rise to unique Lagrangian terms at each order in $\pi$ up to total derivatives and arbitrary coefficients. Remarkably the highest order, $n$, is determined by the number of dimensions, $d$, of the space-time, where $ n = d+1$. This theory has been extended to include several independent galileon fields \cite{Padilla:2010ir}\cite{Padilla:2010tj}\cite{zhou}\cite{Padilla:2010de}\cite{Hinterbichler:2010xn}, supersymmetry \cite{Khoury:2011da}, curved space generalisations\cite{clare}\cite{trodden} and covariant completion\cite{Deff1}\cite{Deff2}. Although motivated by the DGP model, the theory is interesting and peculiar in its own right. There are novel field theoretic properties both at the classical and quantum mechanical level \cite{Endlich:2010zj}\cite{Padilla:2010ir}. In particular it is possible to choose suitable parameters to avoid ghost instabilities in the self-accelerating branch as opposed to the DGP model where there is no freedom to choose these parameters appropriately \cite{nicolis}. Furthermore violations of null-energy condition can be obtained without causing any instability \cite{Nicolis:2009qm}\cite{Creminelli:2010ba}.
In this paper we derive the Hamiltonian for a single galileon field living in Minkowski background space-time with an arbitrary time-like boundary at spatial infinity. This has previously been done for multi-galileons without taking into account the boundary contribution \cite{zhou}. Here we keep careful track of all the boundary terms and investigate the energy of the static spherically symmetric galileon field at cubic order sourced by a point-mass at the origin. We find that the energies for the non-trivial and normal (Minkowski) branch have equal magnitude but opposite signs depending on the sign of the coefficient of the quadratic term $\alpha_2$ (see (2)). Setting $\alpha_2 > 0$ gives positive (negative) energy for the normal (non-trivial) branch and vice versa, indicating ghost like behaviour in the branch with negative energy as we discuss later. This is a non-linear manifestation of the perturbative ghost instability that has been explored extensively.
Section-1 illustrates the framework used in computing the Hamiltonian. We use the normalization of energy with respect to a reference solution as was done in \cite{Hawking:1995fd}\cite{Padilla:2003qi}. In section-2 we present the ADM 3+1 splitting for the bulk Lagrangian density. In section-3 we do a subsequent decomposition of the boundary terms. We present the general expression for the Hamiltonian in section-4 and in the final section we use this result for a static spherically symmetric single galileon field and explore the implications of this result.
\subsection{Infrared regularization of the Hamiltonian}
Our aim is to calculate the Hamiltonian for single galileon field theory living in Minkowski space-time with closed boundary(see fig1). The boundary is made up of constant-time hypersurfaces at far-past and far-future, $\Sigma_{-\infty}$, $\Sigma_{+\infty}$ and bounded by an arbitrary time-like hypersurface, $B$, at spatial infinity, with no inner boundaries. Usually it is fairly straightforward to calculate the Hamiltonian from the action of a field theory, where the Hamiltonian is the Legendre transformation of the Lagrangian, but it is slightly non-trivial when the action has boundary terms as in GR (Gibbons-Hawking-York boundary term). We follow a method that is conceptually similar to that followed by \cite{Hawking:1995fd} in defining a physically meaningful notion of Hamiltonian for unbounded space-times. This is done by regularizing the action with respect to a reference field as explained below.
The most general action for a single galileon field, $\pi(x)$, in 4-D is given by\cite{nicolis}\cite{Dyer:2009yg},
\begin{equation}
S_{galileon} = S_{bulk} + S_{boundary}
\end{equation}
where,
\begin{equation}
S_{bulk}= \sum_{n=2}^{n=5} \int_{M} L^n
\end{equation}
\begin{equation}
L^n = \left\{ -\alpha_n \, \pi_{a_2}\pi^{[a_2} \pi_{a_3}^{a_3}\dots \pi_{a_n}^{a_n ]} \right\}
\end{equation}
\begin{equation}
S_{boundary} = \sum_{n=3}^{n=5} \int_{\partial M}\left \{ \alpha_n \, (n-2) \pi_{\bot}\,\, \pi_{\tilde a_3} \pi^{[\tilde a_3} \pi_{\tilde a_4}^{\tilde a_4} \dots \pi_{\tilde a_n}^{\tilde a_n]} \right \}
\end{equation}
Here $\pi_{a} = \partial_a \pi$ and $\pi_{\bot},\,\pi_{\tilde a_n}$ are orthogonal and tangential derivatives with respect to the boundary. We use the convention that antisymmetrization over the $a$ indices do not involve the prefactor $\frac{1}{n!}$. Note that in the subsequent sections index, $a$, runs over 0..4 and $i$ runs over 1..3. This is a consistent action for dirichlet boundary condition, ie when the fields and their tangential derivatives are held fixed at the boundary \cite{Dyer:2009yg}. $S_{boundary}$ is the analog of the Gibbons-Hawking-York boundary term in GR. Note that $\alpha_2>0$ as we have defined in (3) yields a stable Minkowski branch free of ghost-like behaviour due to the positivity of the kinetic term, however this would make some other branches unstable. A concrete example of this is infact what we discuss in the final section. The action defined above is finite for compact geometries but diverges for non-compact space-times. To renormalize this action for non-compact space-times we choose a reference background $\pi_0$ that asymptotes to the value of $\pi$ and also a solution of the theory. Then we demand the physical action to be given by,
\begin{equation}
S_{physical} = S_{galileon}[\pi] - S_{galileon}[\pi_0]
\end{equation}
consequently the physical Hamiltonian is,
\begin{equation}
H_{physical} = H_{galileon}[\pi] - H_{galileon}[\pi_0]
\end{equation}
In order to derive the Hamiltonian for galileon field theory, one must do an ADM decomposition of the action. We postpone the final result until we have presented the decompositon of the bulk-space-time and the decomposition of boundary terms in terms of relevant derivatives and geometrical quantities.
\subsection{Bulk Decomposition}
\begin{figure}[h]
\centering
\includegraphics[width=0.5\textwidth]{stime.pdf}
\caption{Space-time with boundary. Here $V^a, U^a$ are vectors orthogonal to hyper-surfaces $B,\Sigma_t$ resp. $r^a,n^a$ are vectors lying on $\Sigma_t,B$ respectively, and orthogonal to $S_t$. }
\label{fig}
\end{figure}
We start with decomposing the action for single galileon field in terms of time and spatial derivatives. This is similar to the ADM formalism in GR except there is a prefered time direction since we are working in Minkowski space-time. We consider the galileon field in Minkowski space-time bounded by a time-like boundary at spatial infinity $B$ (see FIG.\,\ref{fig}). We foliate the bulk space-time in constant-time space-like hypersurfaces $\Sigma_t$. Thus the natural embedding is as follows,
\begin{equation}
\Sigma_t:[x^a] \to [t,x^i]
\end{equation}
where $t,x^i$ define the standard cartesian coordinates giving the line element as,
\begin{equation}
ds^2 = -dt^2 + \delta_{ij}dx^i dx^j
\end{equation}
Ignoring the boundary terms (4), the most general Lagrangian for galileon fields in 4-dimensions can be expressed as follows,
\begin{equation}
L_{galileon} = \sum_{n=2}^{5} L^n
\end{equation}
We consider a general term $L^n$ of order $n$ in $\pi$ and seek to do a 3+1 split in terms of time and space. In the spirit of integrating by parts, we rewrite the Lagrangian as a piece that contains no $2^{nd}$ order time derivatives, $L^n_{bulk}$, and a total derivative term, $L^n_{left-over}$, (see Appendix [I] for details). Thus,
\begin{equation}
L^{n} = L_{bulk}^{n} + L_{left-over}^{n}
\end{equation}
where,
\begin{equation}
L^{(n)}_{bulk} = \alpha_n\left\{ {}^{n}C_{2}\,\,\dot\pi^{2} \pi_{i_3}^{[i_3}\pi^{i_4}_{i_4}....\pi^{i_n]}_{i_n} - \pi_{i_2}\pi^{[i_2}\pi_{i_3}^{i_3}...\pi_{i_n}^{i_n]} \right \} \nonumber
\end{equation}
here,${}^nC_2 = \frac{(n)(n-1)}{2}$.
\begin{align}
L^{n}_{left-over} &= \alpha_n \bigg \{ -\frac{(n-2) (n+1)}{2} \partial_{i_3} \left [\dot\pi^2 \pi^{[i_3}\pi^{i_4}_{i_4}...\pi^{i_n]}_{i_n}\right]-(n-2)\partial^a \left [\pi_a \pi_{i_3} \pi^{[i_3} \pi^{i_4}_{i_4}\dots \pi^{i_n]}_{i_n} \right]
\\\nonumber
&\,\,\,\,\,\,+(n-2)\partial^i\left[\pi_i \pi_{i_3} \pi^{[i_3}\pi^{i_4}_{i_4}... \pi^{i_n}_{i_n} \right]
+(n-2)(n-3)\partial_{i_3} \left[\dot\pi \pi_{i_4} \pi^{[i_3}\pi^{i_4}_{t}\pi^{i_5}_{i_5}...\pi^{i_n]}_{i_n} \right] \bigg \}
\end{align}
Inserting the boundary term (4) back into the action and using Stoke's Theorem to convert bulk-integrals to boundary-integrals and including terms of all-order in 4D we recast the total action as follows,
\begin{equation}
S_{total} = S_{bulk} + S_{total-boundary}
\end{equation}
where
\begin{equation}
S_{bulk} = \sum_{n=2}^5 \int dt \int_{\Sigma_t} \alpha_n \left\{ {}^{n}C_{2}\,\,\dot\pi^{2} \pi_{i_3}^{[i_3}\pi^{i_4}_{i_4}....\pi^{i_n]}_{i_n} - \pi_{i_2}\pi^{[i_2}\pi_{i_3}^{i_3}...\pi_{i_n}^{i_n]} \right \}
\end{equation}
\begin{equation}
S_{total-boundary} = \sum_{n=3}^5 S^n_{total-boundary}
\end{equation}
with,
\begin{align}
S^{n}_{total-boundary}&=\alpha_n \int dt \int_{S_t}\left \{ -\frac{(n-2)(n+1)}{2} r_{i_3}\left [ \dot \pi^2 \pi^{[i_3}\pi_{i_4}\pi^{i_4} \dots \pi_{i_n}^{i_n]} \right] + (n-2)\pi_r \left [ \pi_{i_3}\pi^{[i_3} \pi_{i_4}^{i_4} \dots \pi_{i_n}^{i_n]} \right] \right. \\\nonumber
&\,\,\,\,\,\,+ (n-2)(n-3) \dot \pi \pi_{i_4} r_{i_3} \pi^{[i_3}\dot \pi^{i_4} \pi^{i_5}_{i_5} \dots \pi_{i_n}^{i_n]} \bigg \}\\\nonumber
&\,\,\,\,\,\,+\alpha_n \int_{\partial M} \left \{(n-2)\pi_V \left [ \pi_{\bar a_3} \pi^{[\bar a_3} \pi_{\bar a_4}^{\bar a_4} \dots \pi_{\bar a_n}^{\bar a_n ]} - \pi_{i_3}\pi^{[i_3} \pi_{i_4}^{i_4} \dots \pi_{i_n}^{i_n]}\right ] \right \}
\end{align}
Note that $S^2_{total-boundary} = 0$. Here $\pi_V = V^a\partial_a \pi, \,\pi_r = r^a \partial_a \pi$ denote the derivatives along the normal vectors $V^a, r^a$ (see FIG.\,\ref{fig}) respectively. $\pi_{\bar a}$ denotes the covariant derivative with respect to the boundary, $B$ (see next section). Also, $\int_{\partial M} = \int dt N \int_{S_t}$ where $ N = (1 + \theta^2)^{-1/2} , \theta = n_ar^a $ is the lapse function. We have completed the decomposition of the bulk-terms in the action. In the next section we decompose the boundary term $S_{total-boundary}$ with respect to the relevant derivatives to be defined below.
\subsection{Boundary decomposition}
We seek to decompose the boundary terms in terms of derivatives with respect to the closed 2-surface $S_t = B \cap \Sigma_t $ and derivatives along $r^a, U^a$. We work in full-space time coordinates and begin by presenting the definitions of various derivatives and projection operators,
\begin{align}
\gamma_{ab} &= g_{ab} + U_aU_b := \textrm{Projection operator for} \, \Sigma_t\\
H_{ab} &= g_{ab} - V_aV_b := \textrm{Projection operator for}\, B\\\nonumber
q_{ab} = H_{ab} + n_a n_b &= \gamma_{ab} - r_a r_b := \textrm{Projection operator for}\, S_t
\end{align}
We use $D_a, \bar D_a , \hat D_a$ to denote covariant-derivatives with respect to $\Sigma_t, B, S_t $. For brevity this convention is used on the indices in long expressions. $\pi_n, \pi_V, \pi_r, \dot \pi$ are derivatives along the corresponding vector fields defined as $\pi_n := D_n \pi := n^a \bar D_a \pi$ etc. Also, $\pi_{n\hat a} := \hat D_a D_n \pi, \,\,\pi_{r \hat a} := \hat D_a D_r \pi, \,\,\pi_{n^2} := D_n^2\pi := D_nD_n\pi, \,\,\pi_{r^2}:= D_r^2 \pi := D_rD_r\pi$. The action of a covariant derivative $\tilde D_a$ on a hypersurface (with an associated projection tensor $h_{ab}$) on a given tensor lying on the surface is given by \cite{wald},
\begin{equation}
\tilde D_a T^{b_1\dots b_i}_{c_1 \dots c_j} = h_a^b h^{b_1}_{d_1} \dots h^{b_i}_{d_i} h_{c_1}^{e_1} \dots h_{c_j}^{e_j} \nabla_b T^{d_1 \dots d_i} _{e_1 \dots e_j}
\end{equation}
Boundary terms contain derivatives $ D_a, \bar D_a, D_a D_b, \bar D_a\bar D_b $ which can be decomposed as follows (see Appendix [II]).
\begin{align}
D_a\pi &= \hat D_a \pi+ r_a D_r\pi\\\nonumber
\bar D_a\pi &= \hat D_a\pi - n_a D_n\pi\\\nonumber
D_a D_b \pi &= \hat D_a \hat D_b \pi + K^{1}_{ab} D_r \pi + 2 r_{(a} \hat D_{b)} D_r \pi - 2r_{(a} K^{1}_{b)c} \hat D^c \pi + r_ar_b D_r^2 \pi \\\nonumber
\bar D_a \bar D_b \pi &= \hat D_a \hat D_b \pi - K^{2}_{ab} D_n \pi - 2 n_{(a}\hat D_{b)} D_n \pi + 2 n_{(a}K^{2}_{b)c} \hat D^c \pi + n_an_b D_n^2 \pi
\end{align}
Here $K^1_{ab}, K^2_{ab} = \theta K^1_{ab}$ are extrinsic curvatures of the 2-surface $S_t$ with respect to the hypersurfaces $\Sigma_t, B$ respectively. $T_{(ab)} = \frac{1}{2}(T_{ab} + T_{ba})$ denotes symmetrization of the indices. We can now express the boundary terms given in (16) by substituting the decomposition given above. Thus the boundary terms are (here we omit the result for $5^{th}$ order for brevity, see Appendix [II]),
\begin{align}\label{eq:s3}
S_{total-boundary}^3 &= \alpha_3 \int dt \int _{S_t} \bigg \{ -3 (1+\theta^2) \dot \pi ^2 \pi_r - \theta^2 \pi_r^3 - 3 \theta(1+\theta^2)^{\frac{1}{2}} \dot \pi \pi_r^2 - \theta (1+\theta^2)^{1/2} \dot \pi^3 + (\hat D \pi)^2 \pi_r \bigg \} \\
&{}\nonumber
\end{align}
\begin{align}\label{eq:s4}
S_{total-boundary}^4 &= \alpha_4 \int dt \int _{S_t}\bigg \{ -2 {\pi_{\hat a}} {\pi^{[\hat a} } K^{1b]}_b \left [ \theta^2 \dot \pi^2 + 2 \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi {\pi_r} + {(1+\theta^2)} {\pi_r}^2 \right]
\\\nonumber
&\,\,\,\,\,\,-2(1 + \theta^2)^{-\frac{1}{2}} {(\hat D \pi)^2 } \bigg[ \theta {(1+\theta^2)} \dot \pi \ddot \pi + 2 \theta^2 {(1+\theta^2)^{\frac{1}{2}}} \dot \pi \dot\pi_r + \theta {(1+\theta^2)} \dot \pi \pi_{r^2} + (1+\theta^2)^{\frac{3}{2}} \pi_r \ddot \pi
\\\nonumber
&\,\,\,\,\,\,+ 2 \theta {(1+\theta^2)} \dot \pi_r {\pi_r} + (1+\theta^2)^{\frac{3}{2}} {\pi_r} \pi_{r^2} - K^B_{nn}\left ( \theta^2 \dot \pi^2 + {(1+\theta^2)} \pi_r^2 + 2 \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi \pi_r \right) \bigg]
\\\nonumber
&\,\,\,\,\,\,+4 \bigg[ \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi^2 {\pi_{\hat a}} \dot\pi^{\hat a} + {(1+\theta^2)} {\pi_r}^2 {\pi_{\hat a}} {\pi_r^{\hat a}} + 2\theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi {\pi_r} {\pi_{\hat a}} {\pi_r^{\hat a}} + (1+2\theta^2)\dot \pi {\pi_r} {\pi_{\hat a}} \dot \pi^{\hat a} + \theta^2 \dot \pi^2 {\pi_{\hat a}} {\pi_r^{\hat a}}
\\\nonumber
&\,\,\,\,\,\,+\theta {(1+\theta^2)^{\frac{1}{2}}} {\pi_r}^2 {\pi_{\hat a}} \dot\pi^{\hat a} \bigg] - 4 \bigg [ {\pi_{\hat a}} {\pi_{\hat b}} K^{1ab} \left [ \theta^2 \dot \pi^2 + {(1+\theta^2)} \pi_r^2 + 2 \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi {\pi_r} \right ]\bigg ]
\\\nonumber
&\,\,\,\,\,\,-2 {(\hat D^2 \pi)} \bigg[ \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi^3 + {(1+\theta^2)} {\pi_r}^3 +(1+3\theta^2) \dot \pi^2 {\pi_r} + 3 \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi {\pi_r}^2 \bigg ]
\\\nonumber
&\,\,+2{(1+\theta^2)}^{-\frac{1}{2}}(\theta \dot \pi \! + \! {(1+\theta^2)^{\frac{1}{2}}} \pi_r) K^1 \bigg [ \theta {(1+\theta^2)}^{\frac{3}{2}} \dot \pi^3 \!+\! 3 \theta^2 {(1+\theta^2)} \dot \pi^2 \pi_r \!+\! 3 \theta^3 {(1+\theta^2)^{\frac{1}{2}}} {\pi_r}^2 \dot \pi \!+\! (\theta^4 - 1) {\pi_r}^3 \bigg ]
\\\nonumber
&\,\,\,\,\,\,-5\dot \pi^2 {\pi_r}^2 K^1 - 6 \dot\pi^{2} {(\hat D^2 \pi)} {\pi_r} - 5 \dot \pi^2 K^1_{ab} {\pi^{\hat a}} {\pi^{\hat b}} + 4 \dot \pi^2 {\pi_{r\hat a}} {\pi^{\hat a}} - 4 {\pi_r}^2 {\pi_{r\hat a}} {\pi^{\hat a}} + 4 {\pi_r}^2 K^1_{ab} {\pi^{\hat a}} {\pi^{\hat b}}
\\\nonumber
&\,\,\,\,\,\,+2{\pi_r}^3 {(\hat D^2 \pi)} + 2{\pi_r}^4 K^1 + 2 {\pi_r} {\pi_{\hat a}} \pi^{[\hat a} \pi^{\hat b ]}_{\hat b} + 2 {\pi_r}^2 \pi_{\hat a} \pi^{[\hat a} K^{1 \hat b]}_{\hat b} + 2 {\pi_r} \pi_{r^2} {(\hat D \pi)^2 } - 2 \dot \pi \dot \pi_r {(\hat D \pi)^2 } \bigg\}
\end{align}
\subsection{Derivation of the Hamiltonian}
Having recast the galileon action interms of ADM decomposition we can now write down the Hamiltonian directly. The Hamiltoninan density of the galileon theory described by the Lagrangian $L(\pi,\partial \pi ,\partial \partial \pi)$ is given by the legendre transform,
\begin{equation}
H = p \dot \pi - L
\end{equation}
where the canonincal momenta p is given by,
\begin{equation}
p = \frac{\partial L}{\partial \dot \pi}
\end{equation}
Thus the Hamiltonian of single galileon field theory is,
\begin{align}
H_{galileon} &= \bigg\{ \sum_{n=2}^{5}\alpha_n\int_{\Sigma_t} \bigg [{}^nC_2\,\,\dot\pi^2\pi_{i_3}^{[i_3} \pi_{i_4}^{i_4} \dots \pi_{i_n}^{i_n]} + \pi_{i_2}\pi^{[i_2} \pi_{i_3}^{i_3} \dots \pi_{i_n}^{i_n]} \bigg] \bigg \}\\\nonumber
&- S^3_{total-boundary} - S_{total-boundary}^4 - S_{total-boundary}^5
\end{align}
Where the last 3 boundary-terms are given by \eqref{eq:s3},\eqref{eq:s4} and \eqref{eq:s5} (Appendix [II]).
\subsection{Energy of static galileon fields coupled to a point-source}
Let us use our Hamiltonian to compute the energy of a single galileon field at cubic order in a static configuration with $SO(3)$ symmetry, coupled to a point mass, $m$, at the origin. We take $S_t$ to be a 2-sphere with fixed radius R. Here the theory contains two vacua: a normal branch ($X^{ref}_+$) and a non-trivial branch ($X^{ref}_-$) (see \eqref{eq:refsol}). The stability of these branches depends on the sign of $\alpha_2$, where $\alpha_2 >0$ leads to a stable normal branch but an unstable non-trivial branch and vice versa. Here we demonstrate that this perturbative instability is consistent with our non-linear calculation using the full Hamiltonian where it manifests as negative energy for non-trival (normal) branch when $\alpha_2$ is positive (negative). The natural coordinates to work with are the spherical coordinates ($r,\theta,\phi$). The Hamiltonian function for this set-up becomes,
\begin{equation} \label{eq:h}
H = 4\pi \int dr r^2 \left\{ \alpha_2 \pi'^2 + 2\alpha_3 \frac{\pi'^3}{r} + \frac{\rho}{M_p} \pi \right \}
\end{equation}
Here $M_p$ is a dimension-full coupling constant with mass dimension, usually this is of order planck mass for gravitational theories. Also, $(')= \frac{d}{dr}$ and $\rho = m\delta^{(3)}(r)$. Note that $S^3_{total-boundary}$ vanishes for this set up, since for static $SO(3)$ symmetric galileon field, $\dot \pi = \hat D_a \pi = 0$, and time invariance of the 3-boundary, $B$, implies $\theta :=n.r = 0$. The equation of motion is given by \cite{nicolis},(The $\pi$ appearing in the expressions from (26) to (33) is the number $\pi$ not to be confused with the field.)
\begin{equation} \label{eq:eom}
\alpha_2 X + 3 \alpha_3 X^2 = \frac{m}{8M_p\pi r^3}
\end{equation}
where $X = \frac{\pi'}{r}$. The normal and non-trivial branch solutions of \eqref{eq:eom} are given implicitly by,
\begin{align}\label{eq:sol}
X_+ &:= \frac{\pi'_+}{r} = \frac{-\alpha_2 + \sqrt{\alpha_2^2 + \frac{3m\alpha_3}{2M_p\pi r^3}}}{6 \alpha_3}\\\nonumber
X_- &:= \frac{\pi'_-}{r} = \frac{-\alpha_2 - \sqrt{\alpha_2^2 + \frac{3m\alpha_3}{2M_p\pi r^3}}}{6 \alpha_3}
\end{align}
The corresponding reference solutions which we choose to be the normal and non-trivial vacuum solutions are given by setting $m=0$ in \eqref{eq:sol}.
\begin{align} \label{eq:refsol}
X^{ref}_+ &= 0\\\nonumber
X^{ref}_- &= -\frac{\alpha_2}{3\alpha_3}
\end{align}
It is convenient to rewrite the integrand in \eqref{eq:h} using the equation of motion to eliminate the $\pi$ dependence. Thus,
\begin{equation}
H = -4 \pi \int_0^R dr r^4 \left\{ \alpha X^2 + 4 \alpha_3 X^3 \right \} + \frac{m}{M_p} \int _0^R dr \left \{ r X \right \}
\end{equation}
The energy for positive and negative branches is now given by,
\begin{equation}
E_\pm = H[X_\pm] - H[X^{ref}_\pm]|_{m=0}
\end{equation}
Substituting \eqref{eq:sol}, \eqref{eq:refsol} above we get,
\begin{align}
E_+ = - E_- &= \frac{2\pi \alpha_2^3 R^5}{135 \alpha_3^2} + \frac{\alpha_2 m}{18M_p \alpha_3} \int_0^R dr r \left(1 + \frac{3\alpha_3 m}{2M_p\pi \alpha_2^2} \,r^{-3}\right)^{1/2}
\\\nonumber
&- \frac{2\alpha_2^3 \pi}{27 \alpha_3^2} \int_0^R dr r^4 \left( 1 + \frac{3\alpha_3m}{2M_p\pi\alpha_2^2}\,r^{-3}\right )^{1/2}
\end{align}
After some change of variables the integrals can be recognized as a linear combination of hypergeometric functions given by,
\begin{align} \label{eq:energy}
E_+ &= \frac{2\pi \alpha_2^3 R^5}{135 \alpha_3^2} + \frac{ sign(\alpha_2) (\frac{m}{M_p})^{3/2} \sqrt{R}}{3\sqrt{6\pi\alpha_3}} {}_2F_1 \left[ -1/2, 1/6,7/6,-\frac{2 M_p\alpha_2^2 \pi R^3 }{3\alpha_3 m}\right ]
\\\nonumber
&- sign (\alpha_2)\frac{2 \alpha_2^2}{63} \sqrt{\frac{2\pi m}{3M_p\alpha_3^3}}R^{\frac{7}{2}}{}_2F_1\left [ -1/2,7/6,13/6,-\frac{2 M_p\alpha_2^2 \pi R^3 }{3\alpha_3 m}\right]
\end{align}
Here the hypergeometric functions are real and positive and defined for the range $\alpha_3 > - \frac{2 M_p \alpha_2^2 \pi R^3}{3 m}$. However for real values of $E_+,E_-$, $\alpha_3$ is forced to be positive.We now take the limit $R \to \infty$ and the energy becomes,
\begin{align}
E_+^{\infty} = -E_-^{\infty} =-\left(\frac{2}{3}\right)^{\frac{7}{3}} \Gamma\left(-\frac{8}{3}\right)\Gamma\left(\frac{7}{6}\right) \frac{(\alpha_2 \alpha_3)^{-\frac{1}{3}}(\frac{m}{M_p})^{\frac{5}{3}}}{\pi^{\frac{7}{6}}}>0
\end{align}
We get a finite expression for energy with equal magnitiude and opposite sign. The infra-red divergence is regularized by substracting the vacuum energy contribution. As a non-trivial check for our calculation we take the limit $m \to 0$ in \eqref{eq:energy} and obtain,
\begin{equation}
\lim_{m \to 0} E_\pm = 0
\end{equation}
as expected.
\subsection{Discussion}
We conclude with a few remarks on our analysis of the energy of galileon field theory. Having presented the expression for Hamiltonian in ADM formalism carefully keeping track of all the boundary terms, we calculated the energy of static spherically symmetric configuration. In particular, the results of our calculation shows,
\begin{itemize}
\item The two branches of the cubic theory coupled to a point source have energies of equal magnitude and opposite sign.
\item The expression for energy flips sign when the sign of $\alpha_2$ is changed.
\item Even though we couple galileon field to a divergent source at the origin, energy is still finite where non-linear cubic contribution dominates the divergent quadratic term and regularizes it.
\end{itemize}
We argue that the negative energy of the non-trival(normal) branch when $\alpha_2 > 0(< 0)$ with a coupling to a point mass indicates a ghost like instability. Our calculations have been entirely classical and as was argued in \cite{Cline:2003gs} the appearance of negative energy can be traced back to the wrong sign in the propagator, at quantum level. If one evades negative probabilities by shifting the poles in the denominator of the propagator it leads to negative energy. Scattering processes involving ghost like particles and ordinary matter particles can generate ghost particles with unbounded negative energy and matter particles with unbounded positive energy. We believe the sign flip of the energy when changing the sign of $\alpha_2$ further reinforces this argument, for it is the correct sign of $\alpha_2$ in ordinary field theories that ensures the positivity of the kinetic term in the Lagrangian. It is interesting to note that a similar calculation was done for Gauss-Bonnet gravity in \cite{Padilla:2003qi} and the authors found that the energies for the 2-branches match both in magnitude and sign. Further more it was shown that one of the vacua of Gauss-Bonnet gravity was unstable despite the fact that ghost like modes were not excited by the spherically symmetric black-hole\cite{tony+charm}. In contrast here we find that point source which can be taken to be a spherically symmetric source in the limiting case, does seem to excite ghost-like modes giving negative energy. We would like to pursue this line of enquiry in future, it would be interesting to do this calculation for covariant galileon model and Multi-galileon theories. It is well known that the bulk part of the Hamiltonian vanishes identically for diffeomorphism invariant field theories \cite{Andrade:2010hx} and it is not clear how this would play out for galileon models. In order to understand the origin of this negative energy it might be interesting to study the instanton transition amplitudes via bubble nucleation between the different vaccua in galileon models using the method pioneered by S.Coleman \cite{coleman}.
\begin{acknowledgements}
I am deeply grateful to my supervisor Antonio Padilla without whose valuable and helpful suggestions this work would not have been possible. I would also like to thank him for reading through the manuscript carefully.
\end{acknowledgements}
\subsection{Appendix I - Bulk Decomposition in Detail}
Consider the general $n^{th}$ order term appearing in the $\pi-Lagrangian$. The highest order possible is $(n+1)$ in space-time of n-dimensions.
\begin{equation}
L^{(n)} = -\alpha_n \left\{ \pi_{a_2}\pi^{[a_2} \pi^{a_3}_{a_3} \dots \pi^{a_n]}_{a_n} \right \}
\end{equation}
First we make note of the following general identities which we would make use of repeatedly. Note that Einstein-summation is assumed for repeated indices.
\begin{equation}\label{eq:split}
T^{a_1 a_2 \dots a_n}_{a_1 a_2 \dots a_n} = T^{t i_2 \dots i_n}_{t i_2 \dots i_n}+ \dots + T^{i_1 i_2 \dots t \dots i_n}_{i_1 i_2 \dots t \dots i_n}+ \dots + T^{i_1 i_2 \dots i_{n-1} t}_{i_1 i_2 \dots i_{n-1} t} + T^{i_1 i_2 \dots i_n}_{i_1 i_2 \dots i_n}
\end{equation}
\begin{equation}\label{eq:anti}
\pi^{[t}\pi^{i_1}_{i_1} \pi^{i_2}_{i_2} \dots \pi^{i_n]}_{i_n} = \pi^{t}\pi^{[i_1}_{i_1}\pi^{i_2}_{i_2} \dots \pi^{i_n]}_{i_n} - n\pi^{i_1}\pi^{[t}_{i_1}\pi^{i_2}_{i_2}\dots \pi^{i_n]}_{i_n}
\end{equation}
\begin{equation}\label{eq:tderiv}
T^{[t,i_1,i_2 \dots ,i_n]}_{ i_1,i_2 \dots ,i_n} = T^{t[i_1 i_2 \dots i_n] }_{i_1 i_2 \dots i_n} - \frac{1}{(n-1)!} T^{i_1[t i_2 \dots i_n]}_{[i_1 i_2 \dots i_n]}
\end{equation}
Using \eqref{eq:split} $L^n$ can be cast in the following form.
\begin{equation}
L^{(n)} = -\alpha_n \left \{ \pi_{t}\pi^{[t} \pi^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n} + (n-2) \pi_{i_2}\pi^{[i_2} \pi^{t}_{t} \dots \pi^{i_n]}_{i_n} + \pi_{i_2}\pi^{[i_{2}} \pi ^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n}\right \}
\end{equation}
Integrating by parts with respect to the upper time index in the second term before doing the anti-commutation operation we get,
\begin{align}
L^{(n)}&= -\alpha_n \bigg \{ (n-1) \pi_{t}\pi^{[t} \pi^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n}+ \pi_{i_2}\pi^{[i_{2}} \pi ^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n} + (n-2) \partial^{[t |} \left[ \pi_t \pi_{i_3} \pi^{|i_3} \pi_{i_4}^{i_4} \dots \pi_{i_n}^{i_n]} \right ]\bigg \}
\\\nonumber
&\textrm{using \eqref{eq:anti} on the first term and \eqref{eq:tderiv} on the last term we get,}
\\\nonumber
\\\nonumber
&=-\alpha_n \bigg\{ (n-1) \pi_{t}\pi^{t} \pi^{[i_3}_{i_3}\dots \pi^{i_n]}_{i_n} - (n-1)(n-2) \pi_t \pi^{i_3} \pi^{[t}_{i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n]}_{i_n} +\pi_{i_2}\pi^{[i_{2}} \pi ^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n}
\\\nonumber
&\,\,\,\,\,\,+ (n-2) \partial^t \left[ \pi_t \pi_{i_3} \pi^{[i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n]}_{i_n} \right ] -\frac{(n-2)}{(n-3)!} \partial^{i_3} \left[ \pi_t \pi^{[t} \pi_{[i_3 } \pi_{i_4}^{i_4} \dots \pi^{i_n]}_{i_n} \right] \bigg \}
\\\nonumber
&\textrm{again using \eqref{eq:anti} for the last term we get,}
\\\nonumber
\\\nonumber
&=-\alpha_n \bigg\{ (n-1) \pi_{t}\pi^{t} \pi^{[i_3}_{i_3}\dots \pi^{i_n]}_{i_n} - (n-1)(n-2) \pi_t \pi^{i_3} \pi^{[t}_{i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n]}_{i_n}+ (n-2) \partial^t \left[ \pi_t \pi_{i_3} \pi^{[i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n]}_{i_n} \right ]
\\\nonumber
&\,\,\,\,\,\,-\frac{(n-2)}{(n-3)!} \partial^{i_3} \left[ \pi_t \pi^t\pi_{[i_3 } \pi^{[i_4}_{i_4} \dots \pi^{i_n]}_{i_n} \right] + \frac{(n-2)(n-3)}{(n-3)!} \partial^{i_3} \left[ \pi_t \pi^{i_4}\pi_{[i_3} \pi^{[t}_{i_4} \pi^{i_5}_{i_5} \dots \pi^{i_n]}_{i_n ]} \right ] + \pi_{i_2}\pi^{[i_{2}} \pi ^{i_3}_{i_3} \dots \pi^{i_n]}_{i_n} \bigg\}
\\\nonumber
\\\nonumber
&=\alpha_n \bigg \{ {}^nC_2 \dot \pi^2 \pi^{[i_3}_{i_3} \dots \pi^{i_n]}_{i_n} -\pi_{i_2}\pi^{[i_2}\pi^{i_3}_{i_3}\dots \pi^{i_n]}_{i_n} - \frac{(n-2)(n+1)}{2} \partial_{i_3} \left[ \dot \pi^2 \pi^{[i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n}_{i_n}\right]
\\\nonumber
& -(n-2)\partial^t\left[ \dot \pi \pi_{i_3}\pi^{[i_3} \pi_{i_4}^{i_4} \dots \pi_{i_n}^{i_n]} \right] + (n-2)(n-3) \partial_{i_3}\left[\dot\pi \pi_{i_4}\pi^{[i_3}\pi_t^{i_4} \pi^{i_5}_{i_5} \dots \pi^{i_n]}_{i_n}\right] \bigg \}
\end{align}
In the final step we have recast the following term as,
\begin{equation}
\pi_t \pi^{i_3} \pi^{[t}_{i_3} \pi^{i_4}_{i_4} \dots \pi^{i_n]}_{i_n} = \frac{1}{2}\pi^{i_3} \left(\pi_t \pi^{t}\right)_{[,i_3|}\left[ \pi^{i_4}_{|i_4} \dots \pi^{i_n}_{i_n]}\right]
\end{equation}
and integrated by parts with respect $i_3$ inside the commutator. Now we convert the term involving $\partial^t\left[..\right]$ into a total derivative in full space-time by adding and substracting a corresponding term involving a total derivative with respect to the spatial slices $\Sigma_t$. Thus we get,
\begin{align}
L_n &= \alpha_n \bigg\{ {}^nC_2 \dot \pi^2 \pi^{[a_3}_{a_3} \dots \pi^{a_n]}_{a_n} - \pi_{a_2}\pi^{[a_2} \pi^{a_3}_{a_3} \dots \pi^{a_n]}_{a_n} -\frac{(n-1)(n+1)}{2} \partial_{a_3} \left[\dot \pi^2 \pi^{[a_3} \pi^{a_4}_{a_4} \dots \pi^{a_n]}_{a_n} \right]
\\\nonumber
&\,\,\,\,\,\,- (n-2) \partial^\mu \left[ \pi_\mu \pi_{a_3} \pi^{[a_3} \pi_{a_4}^{a_4} \dots \pi_{a_n}^{a_n]} \right]+ (n-2)\partial^a\left[ \pi_a \pi_{a_3} \pi^{[a_3} \pi^{a_4}_{a_4} \dots \pi^{a_n]}_{a_n} \right]
\\\nonumber
&\,\,\,\,\,\, + (n-2)(n-3) \partial_{a_3} \left[ \dot \pi \pi_{a_4} \pi^{[a_3} \pi_t^{a_4} \pi_{a_5}^{a_5} \dots \pi_{a_n}^{a_n]} \right ] \bigg \}
\end{align}
as promised.
\subsection{Appendix II}
\subsubsection{Decomposing the extrinsic curvature of B}
Extrinsic curvature of the time-like surface B is given by,
\begin{equation}
K^{B}_{ab} = H_a^c \left[\partial_c V_b\right]
\end{equation}
we wish to decompose this interms of the following basis of one forms,
\begin{equation}
E_V = V_a dx^a, \hat E_{a } = q_{ab} dx^b, E_n = n_a dx^a
\end{equation}
we get,
\begin{align}
K^B_{V \hat a} &= V^b q^{ad} K^{B}_{bd} = 0
\\\nonumber
K^B_{VV} &= V^aV^b K^{B}_{ab} = 0
\\\nonumber
K^B_{\hat a \hat b} &= q_a^c q_b^d K^{B}_{cd} = q_a^c q_b^d H_c^e \left[\partial_e V_d\right ] = - q_a^e \left[\partial_eq_b^d\right] V_d = q_a^e \left[\partial_e [r_br^d]\right] V_d = (V.r) K^1_{ab}
\\\nonumber
K^B_{\hat a n} &= q_a ^b n^c K^{B}_{bc} = q_a^b n^c H_b^d \left[\partial_d V_c\right] = q_a^d n^c \left[\partial_d V_c\right]
\\\nonumber
&=q_a^b n^c H_c^d \left[\partial _d V_b\right] = q_a^b n^d \left[\partial_d V_b\right] = V_b n^d \left[\partial_d[r_ar^b]\right] = (V.r) n^d \left[\partial_d r_a\right]
\end{align}
Thus,
\begin{equation}
K^B_{ab} = K_{\hat a \hat b} - 2 \,n_{(a|} K^B_{n |\hat b)} + n_an_b K^B_{nn}
\end{equation}
as expected.
\subsubsection{Decomposing the derivatives $D_aD_b\pi,\,\bar D_a \bar D_b \pi$}
First we derive the following results to be used later,
\begin{align}
&K^1_{ab} = q_a^c D_c r_b = q_a^c \gamma_c^d \gamma_b^e \left[\partial_d r_e\right] = q_a^d \gamma_b^e \left[\partial_d r_e\right] = q_a^d \left[\partial_d[\gamma_b^e r_e]\right] = q_a^d \left[\partial_d r_b \right]
\\\nonumber
\\
&K^2_{ab} = q_a^c \left[\bar D_c n_b\right] = q_a^c H_c^d H_b^e \left[\partial_d n_e\right] = q_a^d H_b^e \left[\partial_d n_e\right] = q_a^d q_b^e \left[\partial_d n_e\right] = q_a^d n_e \left[\partial_d [r_b r_e]\right] = (n.r) q_a^d \left[\partial_d r_b\right] = (n.r)K^1_{ab}
\\\nonumber
\\
&\hat D_a \hat D_b \pi = \hat D_a [q_b^c \,\,\partial_c \pi] = q_a^d q_b^e \partial_d[q_e^c \,\,\partial_c \pi] = q_a^d q_b^c \left[\partial_c \partial_d \pi\right] - q_a^d q_b^e \left[\partial_dr_e\right] D_r \pi = q_a^d q_b^c \left[\partial_c \partial_d \pi\right] - K^1_{ab} D_r \pi
\\\nonumber
\\
&q_a^d r_b r^c \left[\partial_c \partial_d \pi\right] = q_a^d r_b \left[\partial_d(r^c \partial_c\pi)\right] - q_a^d r_b \left[\partial_d r^c\right] \partial _c \pi = r_b \left[\hat D_a D_r \pi\right] - r_b K^1 _{ac} \hat D ^c \pi
\\\nonumber
\\
&r_ar_b r^c r^d \partial_c \partial_d \pi = r_ar_b r^c \partial_c(r^d \partial_d\pi) - r_ar_b r^c (\partial_c r^d) \partial_d \pi = r_ar_b D_r^2 \pi - r_ar_b (r^c D_c r^d) \partial_d \pi = r_ar_b D_r^2 \pi\\\nonumber
& \textrm{we have used the result} \quad r_aD^a r_b = 0 \quad \textrm{in the last equality}.
\\\nonumber
\\
&q_a^d n_b n^c \partial_c \partial_d \pi = n_b q_a^d \partial_d(n^c \partial_c \pi) - n_b q_a^d (\partial_dn^c)(\partial_c\pi) = n_b \hat D_a D_n \pi - n_b q_a^d [H^c_e + V^c V_e] (\partial_d n^e) (\partial_c\pi)
\\\nonumber
&=n_b \hat D_a D_n \pi - n_b K^2 _{ac} \hat D^c \pi - n_b q_a^d V_e \partial_d n^e D_V\pi = n_b \hat D_a D_n \pi - n_b K^2 _{ac} \hat D^c \pi + n_b K^B_{\hat a n} D_V \pi
\\\nonumber
\\
&n_an_bn^cn^d \partial_c\partial_d \pi = n_a n_b n^c \partial_c(n^d \partial_d \pi) - n_an_b (n^c\partial_cn^d) \partial_d\pi = n_an_b D_n^2 \pi - n_an_b [H_e^d + V^d V_e] (n^c \partial_c n^e) \partial_d \pi
\\\nonumber
&=n_an_b D_n^2 \pi - n_an_b (n^c \bar D_c n^d) \partial_d\pi - n_an_b n^c V_e \partial_cn^e D_V \pi = n_an_b D_n^2 \pi + n_an_b K^B_{nn} D_V \pi
\\\nonumber
&\textrm{ we used } \, n_a\bar D^a n_b = 0 \, \textrm{in the last equality.}
\end{align}
\begin{align}
D_aD_b \pi &= D_a[\gamma_b^c \partial_c \pi] = \gamma_a^d \gamma_b^e \partial [ \gamma_e^c \partial_c \pi] = \gamma_a^d \gamma_b^c \partial_c\partial_d \pi
\\\nonumber
&=q_a^d q_b^c \partial_c\partial_d \pi + 2 q_{(a|}^d r_{|b)} r^c \partial_c \partial_d \pi + r_ar_b r^c r^d \partial_c\partial_d \pi
\\\nonumber
&= \hat D_a \hat D_b \pi + K^1 _{ab} + 2 r_{(a|} \hat D_{|b)}D_r \pi - 2 r_{(a|} K^1 _{|b)c} \hat D^c \pi + r_a r_b D_r^2 \pi
\end{align}
where(47),(49),(50), (51) was used.
\begin{align}
\bar D_a \bar D_b \pi &= \bar D_a[H_b^c \partial _c \pi] = H_a^d H_b^e \partial_d [H_e^c \partial_c \pi] = H_a^d H_b^c \partial_c \partial_d \pi +H_a^d H_b^e (\partial_d H_e^c)(\partial_c\pi)
\\\nonumber
&=q_a^d q_b^c \partial_c\partial_d \pi - 2 q^d_{(a|} n_{|b)} n^c \partial_c \partial_d \pi + n_a n_b n^c n^d \partial_c\partial_d \pi - H_a^d H_b^e \partial_d[V_eV^c] \partial_c \pi
\\\nonumber
&=q_a^d q_b^c \partial_c\partial_d \pi - 2 q^d_{(a|} n_{|b)} n^c \partial_c \partial_d \pi + n_a n_b n^c n^d \partial_c\partial_d \pi -K^B_{ab} D_V\pi
\\\nonumber
&=\hat D_a \hat D_b \pi - \theta K^1_{ab} D_n \pi - 2n_{(a|}\hat D_{|b)} D_n \pi + 2 n_{(a|} K^2 _{|b)c} \hat D^c \pi + n_an_b D_n^2 \pi
\end{align}
where (45), (46), (47), (48), (49), (52), (53) was used.
\subsubsection{Boundary term at $5^{th}$ order}
\begin{align}\label{eq:s5}
S^5_{total-boundary} &= \alpha_5 \int dt \int_{S_t} \bigg \{ -9 \dot \pi^2 \bigg [ \pi_{[r} \pi^{\hat b}_{\hat b} \pi^{\hat c}_{\hat c]} + (\pi_r)^3 K^{[1 b}_{b} K^{1 \hat c]}_{\hat c} + 2 \pi_r K^{1[b}_{b} K^{1 c]}_{d} \pi_c \pi^d + (\pi_r)^2 \pi^{[\hat b}_{[\hat b} K^{1 c]}_{c]} + 2 \pi^{[\hat b}_{\hat b} K^{1 c]}_{d} \pi_c \pi^d
\\\nonumber
&-2 \pi_r K^{[1b}_{b}\pi^{\hat c]} \pi_{r\hat c} \bigg ]
\\\nonumber
&+ 3(1+\theta^2)^{-\frac{1}{2}} (\theta \dot \pi + {(1+\theta^2)^{\frac{1}{2}}} \pi_r) \bigg [-(\pi_r \pi_{[r|} + \pi_n \pi_{[n|})\pi^{\hat b}_{|\hat b}\pi^{\hat c}_{\hat c]} -(\theta^2 \pi_n^4 + \pi_r^4) K^{1[\hat b}_{\hat b}K^{1 \hat c]}_{\hat c}
\\\nonumber
&+ 2(\theta^2 \pi_n^2 - \pi_r^2) \pi_{\hat c} \pi^{\hat d} K^{[1b}_b K^{1c]}_d +(\theta \pi_n^3 - \pi_r^3) \pi^{[\hat b}_{[\hat b} K^{1 c]}_{c]} - 2(\theta \pi_n^2 \pi_{n\hat c} - \pi_r^2 \pi_{rc}) K^{[1\hat b}_{\hat b} \pi^{\hat c]}
\\\nonumber
&- 2(\theta \pi_n + \pi_r) \pi^{[\hat b}_{\hat b} K^{1c]}_{d}\pi_{\hat c} \pi^{\hat d} + (\theta^2 \pi_n^2 - \pi_r^2) \pi_{\hat a} \pi^{[\hat a} K^{1b}_{b} K^{1 c]}_{c} + 2 \pi_{\hat a} \pi^{[\hat a} \pi_n^{\hat b ]}\pi_{n \hat b} + 2 \pi_{\hat a} \pi^{[\hat a} \pi_r^{b]}\pi_{r \hat b}
\\\nonumber
&-2\theta \pi_{\hat a} \pi^{[\hat a}K^{1b]}_{c} \pi_{n\hat b}\pi^{\hat c} - 2 \pi_{\hat a} \pi^{[\hat a}K^{1 b]}_{c} \pi_{r \hat b} \pi^{\hat c} - 2 \theta K^{1c}_b \pi_{\hat a} \pi^{[\hat a}\pi_n^{\hat b]}\pi_{\hat c}
-2 K^{1b}_{c}\pi_{\hat b} \pi_{\hat a} \pi_r^{[\hat c} \pi^{\hat a ]}
\\\nonumber
&+2\theta^2 \pi_{\hat a} \pi^{[\hat a} K^{1 \hat b]}_{c} K^1_{bd} \pi^{\hat c} \pi^{\hat d} + 2\pi_{\hat a} \pi^{[\hat a} K^{1b]}_{c} K^1_{bd} \pi^{\hat c} \pi^{\hat d} - (\theta \pi_n + \pi_r)\pi_{\hat a} \pi^{[\hat a}\pi^{\hat b}_{[\hat b} K^{1 c]}_{c]} - 2(\pi_{n^2} + \pi_{r^2}) \pi^{[\hat a} \pi_{\hat b}^{\hat b]}\pi_{\hat a}
\\\nonumber
&+2(\theta \pi_n \pi_{n^2} - \pi_r \pi_{r^2}) \pi_{\hat a} \pi^{[\hat a} K^{1\hat b]}_{\hat b} - 2 \theta \pi_n^2 \pi_{\hat a} \pi_n^{[\hat a} K^{1b]}_{b} + 2 \pi_r^2 \pi_{\hat a} \pi_r^{[\hat a }K^{1b]}_{b} + 2 (\theta^2 \pi_n^2 - \pi_r^2) \pi_{\hat a} K^{1[a}_{c}K^{1b]}_b \pi^{\hat c}
\\\nonumber
&+2\pi_n \pi_{\hat a} \pi_n^{[\hat a} \pi^{\hat b]}_{\hat b} + 2 \pi_r \pi_{\hat a} \pi_r^{[\hat a} \pi_{\hat b}^{\hat b ]} - 2(\theta \pi_n + \pi_r) \pi^{[\hat b}_{\hat b} K^{1 a]}_{c} \pi^{\hat c} \pi_{\hat a} \bigg]
\\\nonumber
&+3 \bigg[\pi_r^2 \pi_{[r}\pi_{\hat b}^{\hat b}\pi ^{\hat c}_{\hat c]} +\pi_r^5 K^{[1b}_{b} K^{1c]}_c + 2\pi_r^3 K^{[1b}_b K^{1c]}_{d}\pi_{\hat c} \pi^{\hat d} + \pi_r^4 \pi_{[\hat b}^{[\hat b} K^{1c]}_{c]} + 2\pi_r^2 \pi^{[\hat b}_{\hat b} K^{1c]}_{d} \pi_{\hat c} \pi^{\hat d} - 2\pi_r^3 K^{1[b}_{b}\pi^{\hat c]}\pi_{r\hat c}
\\\nonumber
&+ \pi_r \pi_{\hat a} \pi^{[\hat a} \pi_{\hat b}^{\hat b} \pi_{\hat c}^{\hat c]} + \pi_r^3 \pi_{\hat a} \pi^{[ \hat a} K^{1 b}_b K^{1c]}_{c} + 2\pi_r \pi_{r\hat b} \pi_{\hat a} \pi_r^{[\hat a} \pi^{\hat b]} - 2\pi_r \pi_{\hat a} \pi^{[\hat a} K^{1b]}_{c} K^{1}_{bd}\pi^{\hat c} \pi^{\hat d} + 2\pi_r \pi_{\hat a} \pi^{[\hat a}K^{1b]}_c \pi^{\hat c} \pi_{r \hat b}
\\\nonumber
& +2\pi_r K^{1b}_c \pi_{\hat b} \pi_{\hat a} \pi_r^{[\hat c} \pi^{\hat a ]} + \pi_r^2 \pi_{\hat a} \pi^{[\hat a} \pi^{\hat b}_{[\hat b} K^{1 c]}_{c]} +2\pi_{r^2}\pi_r \pi_{\hat a} \pi^{[\hat a} \pi^{\hat b]}_{\hat b} +2\pi_r^2 \pi_{r^2} \pi_{\hat a} \pi^{[\hat a} K^{1b]}_b - 2\pi_r^3 \pi_{\hat a} \pi_r^{[\hat a} K^{1b]}_b
\\\nonumber
&+ 2\pi_r^3 \pi_{\hat a} K^{1[a}_c K^{1b]}_b \pi^{\hat c} - 2\pi_r^2 \pi_{\hat a} \pi_r^{[\hat a} \pi_{\hat b}^{\hat b]} + 2\pi_r^2 \pi_{\hat b}^{[\hat b} K^{1a]}_{c} \pi^{\hat c} \pi_{\hat a} \bigg]
\\\nonumber
&+6 \bigg[ \pi_r \dot \pi \dot \pi_{\hat a} \pi^{[\hat a} \pi_{\hat c}^{\hat c]} + \pi_r^2 \dot \pi \dot \pi_{\hat a} \pi^{[\hat a} K^{1c]}_c - \dot \pi \dot \pi_r \pi_{\hat b} \pi^{[\hat b} \pi^{\hat c]}_{\hat c} - \dot \pi \dot \pi_r \pi_{\hat b} \pi^{[\hat b}K^{1c]}_c \pi_r \bigg] \bigg \}
\end{align}
If one needs to restrict the above expression to the basis $ E_u = U_a dx^a, E_r = r_a dx^a, \hat E_{ a} = q_{ab}dx^b $, the following expressions can be used to convert the relevant terms (we omit this step for brevity),
\begin{align}
&\pi_n = {(1+\theta^2)^{\frac{1}{2}}} \dot \pi + \theta \pi_r
\\\nonumber
&\pi_{n\hat a} = {(1+\theta^2)^{\frac{1}{2}}} \dot \pi_{\hat a} + \theta \pi_{r \hat a } - K^B_{\hat a n} \left [ \theta \dot \pi + {(1+\theta^2)^{\frac{1}{2}}} \pi_r \right]
\\\nonumber
&\pi_{n^2} = {(1+\theta^2)} \ddot \pi + 2 \theta {(1+\theta^2)^{\frac{1}{2}}} \dot \pi_r + \theta^2 \pi_{r^2} - K^B_{nn}\left[ {(1+\theta^2)^{\frac{1}{2}}} \pi_r + \theta \dot \pi \right]
\end{align}
|
\section{Introduction}
In nearly all extensions of the Standard Model, the
generation of neutrino masses leads to the
introduction of sterile neutrinos, e.g., \cite{Dinh:2006ia,Dong:2006mt}.
Sterile neutrinos may also play the role of
the dark matter, e.g., \cite{Dodelson1994,Shi:1998km,AFP,AFT,Dolgov:2000ew,AbaSav,Taoso:2007qk,
Bazzocchi:2008fh, Boyarsky:2009ix, Wu:2009yr, Gelmini:2009xd, Bezrukov:2009th} and
affect a host of interesting cosmological and astrophysical processes, including
the production of baryon \cite{AS2005,ABS2005,Asaka:2006ek,Smith:2008ic} and lepton \cite{ABFW}
asymmetries, Big Bang Nucleosynthesis \cite{Dolgov:2000jw,ABFW,Smith:2008ic}, the evolution of the
matter power spectrum \cite{Boyanovsky:2008nc,Boyanovsky:2010pw}, reionization \cite{
Barkana:2001gr,Yoshida:2003rm,Hansen:2003yj,Gelmini:2004ah,Biermann:2006bu,O'Shea:2006tp,Mapelli:2006ej,
Ripamonti:2006gq,Stasielak:2006br},
neutrino oscillations \cite{Cirelli:2004cz,Smirnov:2006bu},
pulsar kicks \cite{Kusenko:1998bk,Fuller:2003gy,Barkovich:2004jp,Kusenko:2004mm,Fryer:2005sz,Kusenko:2008gh},
and supernovae \cite{Hidaka:2006sg,Hidaka:2007se,Raffelt:2011nc}.
The degree to which sterile neutrinos participate in
these processes is sensitive to the strength of their interactions with Standard Model
particles and their abundance throughout cosmic history. The creation and exploration of
models and production scenarios
that characterize these properties has been
\cite{Dodelson1994,Shi:1998km,AFP,AFT,Dolgov:2000ew}
and continues to be \cite{Aba05a,Aba05b,Shaposhnikov:2006xi,deGouvea:2006gz, Kusenko:2006rh,
Asaka:2006nq,Petraki:2007gq,Khalil:2008kp,
Kadota:2007mv,Lattanzi:2008ds,Kishimoto:2008ic,
Shaposhnikov:2008pf,Laine:2008pg,Jenkins:2008fx,Bazzocchi:2008fh, Boyarsky:2009ix,
Wu:2009yr, Gelmini:2009xd, Bezrukov:2009th} a very active area of research.
See Ref.~\cite{Kusenko:2009up} for a review of sterile neutrino properties.
In addition to their rich phenomenology, sterile neutrinos are also a readily testable
dark matter candidate. To date, their properties have primarily been constrained in
two ways: through X-ray searches for
their radiative decays and via observed cosmological small-scale structure, but the possibility
of detection through atomic ionization and nuclear spin flip signatures has also recently been discussed
\cite{Ando:2010ye}.
Although the two primary methods are generally considered separately,
combined studies of both radiative and cosmological constraints have also been performed
\cite{Palazzo:2007gz,Boyanovsky:2007ay,Bazzocchi:2008fh}.
The radiative decays of predominantly sterile neutrino mass eigenstates to
predominantly active neutrino mass eigenstates and X-rays of energy $E_{\gamma, \rm
s} = m_s/2$ can be detected by existing X-ray satellites \cite{AFT}.
Current radiative decay limits are based on
observations of a long list of sources, including the Cosmic X-ray background
\cite{Boyarsky:2005us,Cumberbatch:2007qq}, nearby clusters such as Virgo \cite{Aba05b} and Coma
\cite{AbaSav,Boyarsky:2006zi}, more distant clusters like A520 and A1835
\cite{RiemerSorensen:2006pi} and the bullet cluster \cite{Boyarsky:2006kc}, nearby galaxies
like Andromeda \cite{Watson:2006, Boyarsky:2007ay}, M33 \cite{Borriello:2011un},
Milky Way satellite galaxies like the Large
Magellenic Cloud \cite{Boyarsky:2006fg}, Ursa Minor \cite{Boyarsky:2006ag, Loewenstein:2008yi},
Draco \cite{RiemerSorensen:2009jp}, Wilman I \cite{Loewenstein:2009cm}, and Seque I \cite{Mirabal:2010an},
as well as unresolved emission from the Milky Way itself
\cite{Boyarsky:2006ag,RiemerSorensen:2006fh,Abazajian:2006jc,Yuksel:2007xh,Boyarsky:2007ge}.
See Refs.~\cite{AbaSav,Watson:2006,Abazajian:2006jc} and references therein for a more detailed summary
and Refs.~\cite{Yuksel:2007dr,deVega:2009ku} for more general studies of radiative limits.
Although the original Dodelson-Widrow (DW) non-resonant sterile neutrino production scenario
\cite{Dodelson1994, AFT, Aba05b} has nearly been excluded by radiative constraints alone, e.g, $m_s < 3.5$ keV
(95\% C.L.) \cite{Watson:2006},
and can account for at most 70\% of the dark matter at 2$_{~}\sigma$ according to
Ref.~\cite{Palazzo:2007gz}, recent Suzaku observations of Wilman I revealed
a spectral feature at 2.5 keV that was consistent with the radiative decay of a 5 keV DW sterile neutrino
\cite{Loewenstein:2009cm}. These findings have been discussed further
in Refs.~\cite{Boyarsky:2010ci,Kusenko:2010sq}, and we will return to
them in Sections~\ref{Analysis}~and~\ref{conclusions} of this paper.
In addition to the
DW scenario, many other models have been proposed that predict the proper
relic dark matter density even at very small active-sterile neutrino mixing angles, e.g.,
\cite{Shi:1998km, AFP, AbaSav, Petraki:2007gq, Laine:2008pg}. General phase space considerations
have also been used to constrain sterile neutrino production scenarios, e.g.,
\cite{Boyarsky:2008ju, Gorbunov:2008ka}.
As a complement to the upper bounds imposed by radiative decay
constraints, observations of the clustering of cosmological structures on the smallest scales can set
lower bounds on the mass of the dark matter particle.
This is because sufficiently light dark matter particles would have suppressed or
even erased these smallest structures by easily propagating (free-streaming) out
of the shallow gravitational potential wells in which they formed.
Current cosmological lower limits on $m_s$ are
based primarily on measurements of small scale clustering in the
CMB, SDSS galaxy, and Ly$\alpha$ forest flux power spectra and gravitational lensing
by the smallest structures. However, the results of N-body
simulations have also recently been used to set lower bounds on $m_s$ based on the requirement that
the number of simulated satellite galaxies equals or exceeds the number of observed Milky Way
satellites \cite{Polisensky:2010rw}.
Of all the cosmological constraints, the Ly$\alpha$ limits are the most sensitive to the
difficult-to-characterize, non-linear growth of baryonic and dark matter structures as well as
the specifics of dark matter production, e.g., the momentum distribution of the
dark matter particles. There have been significant discrepancies between the lower
limits \cite{Aba05b,Seljak:2006qw,Viel:2006kd,Viel:2007mv,Boyarsky:2008xj,Boyarsky:2008mt}
obtained using different data sets and hydrodynamical simulations, as discussed in, e.g.,
\cite{AbaSav,Abazajian:2006jc}.
Ly$\alpha$ limits are less restrictive for
models in which sterile neutrinos behave more like cold dark matter, e.g.,
\cite{AbaSav,Asaka:2006ek,Petraki:2008ef}.
Studies of gravitationally lensed galaxies and QSOs have provided comparable lower limits,
e.g., $m_s > 5.2$ keV \cite{Dalal:2002su} and $m_s >
10$ keV \cite{Miranda:2007rb}, respectively, based on the image distortions caused by the smallest structures.
These lensing constraints should improve significantly as data from future submillilensing
experiments become available \cite{Hisano:2006cj}. Lower bounds based on comparisons of satellite galaxy
counts to N-body simulations,
e.g., $m_s > 13.3$ keV (in the DW scenario) \cite{Polisensky:2010rw}, are also similar to the most restrictive
Ly$\alpha$ constraints.
In this paper, we consider the radiative decay limits imposed by \it{Chandra}\rm~\cite{Weisskopf:2002}~observations
of the Andromeda galaxy (M31). We choose to focus on Andromeda because of 1)
its close proximity, 2) its substantial and well-studied dark matter distribution,
and 3) its intrinsically low level of X-ray emission. These properties tend to maximize the
prospective sterile neutrino decay signal (1 and 2) and minimize noise, i.e., astrophysical background (3).
The \it{Chandra}\rm~data set we use in this study also has several advantages over
the \it{XMM-Newton}\rm~observations of Andromeda \cite{Shirey} that were used to constrain the properties of
sterile neutrinos in Ref.~\cite{Watson:2006} (W06).
First, compared to \it{XMM-Newton}\rm~EPIC, the
\it{Chandra}\rm~ACIS detector has a significantly lower and much more stable
instrumental background, making it particularly well-suited for extracting low surface
brightness, extended X-ray emission.
Second, using \it{Chandra's}\rm~superior angular resolution, we are able
to remove the emission from a larger number of resolved, X-ray point sources than would be possible
with \it{XMM-Newton},\rm~thereby further reducing astrophysical background noise,
while excising very little dark matter from the small excluded regions.
Third, the field of view (FOV) associated with the \it{Chandra}\rm~spectrum
we use in this paper is over 25 times the area of the \it{XMM-Newton}\rm~FOV studied in W06
(a 12 arcminute to 28 arcminute annulus vs.~a circle of 5 arcminute radius) and
therefore probes a significantly larger
dark matter mass and prospective $\nu_s$ decay signal.
The dark matter mass within this FOV is not only significantly larger but
also subject to less uncertainty both because we
exclude the central region of the galaxy, where the dark
matter density profile is least well constrained, and because we make use of an updated model of the
Andromeda dark matter distribution
based on more recent kinematic data and theoretical considerations.
The results of this study are also more accurate because we
use the \it{Chandra}\rm~ACIS-I effective area, rather than an average flux-to-count ratio, to convert
calculated sterile neutrino decay signals into spectral features.
For Majorana sterile neutrinos, the resulting unresolved emission spectrum requires $m_s < 2.2$ keV
(95\% C.L.) to avoid more than doubling the observed signal in bins of energy $E \ge$ 1.1 keV.
Although the mass-mixing exclusion region we generate with this new data set is not vastly more
restrictive than the exclusion region presented in W06 (see Fig. 4 below), it is more robust for the reasons cited above,
and as we argue throughout the paper, it will be difficult to improve upon the constraints we set here with
the current generation of X-ray detectors.
In Sec.~\ref{Model}, we discuss the basics of the DW scenario as well as some more recent
sterile neutrino models.
In Sec.~\ref{M31_prop}, we discuss the \it{Chandra}\rm~observations of Andromeda,
how we generate the unresolved X-ray spectrum we use for our constraints,
and how we calculate the dark matter mass contained within the \it{Chandra}\rm~FOV.
In Sec.~\ref{Analysis}, we present our analysis of the unresolved spectrum, discuss the
limits we are able to impose on the sterile neutrino mass and mixing,
and compare our results to those of previous studies.
In Sec.~\ref{conclusions} we summarize our findings and discuss the outlook
for future sterile neutrino constraints.
Throughout the paper, we assume a flat cosmology
with $\Omega_{\rm baryon}=0.04$, $\Omega_{\rm WDM}= \Omega_{\rm s}= 0.24$,
$\Omega_{\Lambda}=0.72$, and $h = H_0/100~ \hbox{km s}^{-1} ~\rm Mpc^{-1} = 0.72$.
\section{The Sterile Neutrino Warm Dark Matter Model}
\label{Model}
The radiative decay rate for Majorana sterile neutrinos is \cite{PW,Barger}
\begin{equation}
\Gamma_s \simeq 1.36\times 10^{-32} \rm{s}^{-1}
\left(\frac{\sin ^{2}2\theta}{10^{-10}}\right)
\left(\frac{m_s}{\rm~keV}\right)^5.
\end{equation}
The line flux at $ E_{\gamma ,\rm s} = m_s/2$ resulting from the decay of
the fraction of sterile neutrinos that are within the detector field of view (FOV),
$N^{\rm FOV}_s = (M^{\rm FOV}_{\rm DM}/m_s )$ in a halo at a distance $D$
is given by \cite{AFT,Aba05a,Aba05b}
\begin{eqnarray}
\Phi_{\rm x,s}(\sin ^{2}2\theta) = \frac{E_{\gamma ,\rm s}N^{\rm FOV}_{\rm s}\Gamma_s} {4\pi D^2} \simeq
1.0 \times 10^{-17} \rm{erg~cm}^{-2}\rm{s}^{-1} \nonumber \\
\times \left(\frac{M^{\rm FOV}_{\rm DM}}{10^{11}M_\odot }\right)
\left(\frac{D}{\rm Mpc}\right)^{-2}
\left(\frac{\sin ^{2}2\theta}{10^{-10}}\right)
\left(\frac{m_s}{\rm~keV}\right)^5,
\label{Phixs}
\end{eqnarray}
where $\sin ^{2}2\theta$ characterizes the active-sterile neutrino mixing.
Dirac sterile neutrinos would produce only half the flux given by Eqn.~(\ref{Phixs}) \cite{PW,Barger}.
To facilitate comparisons between our work and other results, we will adopt Majorana sterile neutrinos,
which have been more commonly assumed in recent studies,
e.g., \cite{Boyarsky:2007ge,Abazajian:2006jc,Boyarsky:2007ay}\footnote{The decay rate
equation given in Ref.~\cite{Abazajian:2006jc}
is for Majorana rather than Dirac sterile neutrinos.}.
For a QCD phase-transition
temperature of $T_{\rm QCD} = 170$ MeV
and a lepton asymmetry of $L \simeq n_{\rm baryon}/n_\gamma \simeq 10^{-10}$,
the sterile neutrino density-production
relationship for this model is \cite{Aba05a}
\begin{equation}
m_s = 55.5\rm~{keV}\left(\frac{sin^{2}2\theta}{10^{-10}}\right)^{-0.615}
\left(\frac{\Omega_{\rm s}}{0.24}\right)^{0.5}.
\label{Omega_sin2th}
\end{equation}
Combining Eqns.~(\ref{Phixs}) and (\ref{Omega_sin2th}) yields an expression for the line flux
that is independent of
the mixing angle:
\begin{eqnarray}
\Phi_{\rm x,s}(\Omega_{\rm s}) \simeq 7.0 \times 10^{-15} \rm{erg~cm}^{-2}\rm{s}^{-1} \nonumber \\
\times \left(\frac{M^{\rm FOV}_{\rm DM}}{10^{11}M_\odot }\right)
\left(\frac{D}{\rm Mpc}\right)^{-2}
\left(\frac{\Omega_{\rm s}}{0.24}\right)^{0.813}
\left(\frac{m_s}{\rm~keV}\right)^{3.374}.
\label{Phixs_nomix}
\end{eqnarray}
We note that the density-production relationship presented in Ref.~\cite{Asaka:2006nq}
agrees with Eqn.~(\ref{Omega_sin2th}) for sterile neutrino masses ranging from
1 keV $\lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle <}\over{\sim}\ $}} m_{s~} \lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle <}\over{\sim}\ $}} 10$ keV. The sterile neutrino mass limits we determine in
Sec.~\ref{Analysis} are therefore valid for the relationships in both
Refs.~\cite{Aba05a} and \cite{Asaka:2006nq}. It would also
be straightforward to re-evaluate our limits on $m_s$ assuming different
density-production relationships, such as the three resonant production models
associated with large $L$ \cite{AbaSav} or the Shi-Fuller model \cite{Shi:1998km} as
calculated in Ref.~\cite{Laine:2008pg}, all of which are shown in Fig.~\ref{ms_mix} below.
\begin{figure}
\includegraphics[height = .35\textheight, width = .5\textwidth]{FOV.eps}
\caption {Here we show the raw counts associated with the 7 \it{Chandra}\rm~ACIS-I
exposure regions (some of which are overlapping) within a $12'-28'$ annulus about the center
of the Andromeda galaxy. Variations in the brightness among these 7 regions are mainly
due to differences in exposure times (which vary from 5 ks to 20 ks) rather than intrinsic variation
in the emission (see text).
We have also included an ACIS-I image of the central $12'$ to illustrate the extent of the
X-ray emission from hot gas and point sources that we have eliminated by
excising the bulge of the galaxy from our FOV.
\label{ACIS_FOV}}
\end{figure}
\section{Properties of Andromeda}
\label{M31_prop}
\subsection{X-ray Data}
To detect or place the most restrictive constraints on radiatively decaying dark matter,
it is critical to minimize the X-ray background from baryonic sources,
such as diffuse hot gas and X-ray binaries. In M31, both
the diffuse hot gas \cite{Li:2007ud,Li:2009mh}
and
X-ray binaries \cite{Voss:2006az}
exhibit a relatively steep radial distribution toward the galactic
center, suggesting that we can optimize the dark matter signal to
baryonic noise ratio by avoiding the inner regions of the galaxy.
Although most existing {\it{Chandra}} observations of M31 have been aimed toward
the inner bulge, we did find seven observations that were aimed toward the outer
bulge/disk regions and were therefore more well-suited for our purposes.
All seven observations used the Advanced CCD Imaging Spectrometer
(ACIS) I-array as the primary detector, with a $17'$x$17'$ field of view (FOV).
The positions of these observations relative to an annulus of
inner radius 12 arcminutes and outer radius 28 arcminutes ($12' - 28' \simeq 2.8 - 6.4$ kpc at
$D_{\rm M31} \simeq$ $0.78 \pm 0.02$ Mpc \cite{Stanek:1998cu,Jerjen:2003ws})
are shown in Figure~\ref{ACIS_FOV}.
Five of the seven observations (ObsIDs 1576, 1584, 2899, 2901 and
2902) have effective exposure times of $\sim$5 ks, while the other two
observations have longer exposure times of 8 ks (ObsID 7067) and 20 ks
(ObsID 11256), respectively. Each $17'$x$17'$ ACIS-I
region covers 11.4\% of the $12'-28'$ annulus, and
the non-overlapping parts of these 7 pointings collectively
cover $\sim 35\%$. We address sampling
bias later in this section, pointing out that the co-added spectrum from
the selected observations provides, if anything, an overestimate of
the X-ray emission from the entire $12'-28'$ annulus.
We obtained and re-processed the archival data using CIAO version
4.2 and the corresponding calibration files. We followed the
standard procedure of data reduction, e.g., as described in Ref.~\cite{Li:2007ud}.
Briefly, for each observation, we (i) filtered time intervals of
high particle background, (ii) detected and removed discrete sources,
(iii) extracted a spectrum for the unresolved X-ray emission in the
0.5-8 keV range and generated corresponding instrumental response (rmf
and arf) files, and (iv) extracted a fiducial instrumental background
spectrum, using the ``stow background data'' that has been calibrated with
the 10-12 keV count rate.
In a final step, we co-added the
seven spectra and generated exposure-weighted response files, using
the {\sl FTOOL addspec}. Such a procedure is appropriate, since the spectra
were extracted from nearly identical detector regions.
Since the effective area $A_{\rm eff}(E)$ of ACIS is known to decrease gradually with time,
it is important to note that these seven observations were taken over a time span of 9 years.
By using a weighted $A_{\rm eff}(E)$ associated with the co-added spectrum in our analysis below, we are
able to gauge the mean sensitivity of the detector over this time interval.
Additionally, we note that the degradation of the detector mainly
affects energies below 1 keV, and our data set has very little constraining power
at these low energies, as we will show.
It is also noteworthy that the observed flux varies modestly
among the seven observations, within a factor of 1.5 of the
mean values associated with the co-added spectrum. This is a natural
result, since the observations sampled different regions of the
outer bulge and the disk. We note, however, that the sampled regions mostly covered
the southeastern half of the disk -- where the
unresolved X-ray emission is higher than that from the
northeaster side, primarily due to the tilt of M31's
galactic plane \cite{Li:2007ud}. This sampling bias means that the co-added spectrum we use
represents an upper limit on the mean flux
within the $12'-28'$ annulus - precisely what we need
to set a robust upper limit on the mass of the sterile neutrino.
Finally, we note that a correction factor of $(0.114)^{-1}$ has been applied to the
data shown in Fig.~\ref{Galspec} to account for the fact that
the co-added spectrum from the 7 archived observations represents
11.4\% of the projected area of the $12'-28'$ annulus.
\subsection{Dark Matter Enclosed within the \it{Chandra}\rm~\bf{Field of View}}
\label{DMest}
The dark matter halo of M31 has been studied extensively.
Klypin, Zhao and Somerville (KZS) \cite{KZS}, for instance, used a variety of kinematic data to determine its dark
matter density profile, $\rho_{\rm DM, M31}$.
Using more recent data and a more accurate baryonic mass profile, Seigar, Barth and Bullock
(SBB) \cite{Seigar:2006ia} have updated the profile found by KZS.
When we integrate the SBB model of $\rho_{\rm DM, M31}$ over
the volume, $V_{\rm FOV, M31}$, associated
with the projection of the Chandra FOV along the line of sight,
we find that the $12' - 28'$ annulus contains:
\begin{eqnarray}
\Sigma^{\rm FOV}_{\rm DM,M31} \simeq (0.8 \pm 0.08) \times 10^{11} M_\odot \rm{Mpc}^{-2}; \nonumber \\
M^{\rm FOV}_{\rm DM,M31} \simeq (0.49 \pm 0.05) \times 10^{11} M_\odot,
\label{M_DM}
\end{eqnarray}
where
\begin{equation}
\Sigma_{\rm FOV} = \int \frac{\rho_{\rm DM}(|\vec{r}-\vec{D}|)dV_{\rm FOV}}{ r^{2}}
\label{Sigma}
\end{equation}
is the dark matter column density within $V_{\rm FOV}$,
and the dark matter mass within $V_{\rm FOV}$ is defined by $M^{\rm FOV}_{\rm DM} = D^{2}\Sigma^{\rm FOV}_{\rm DM}$.
(See W06 for further details).
In addition to being based on more contemporary data and theoretical considerations, the SBB value
for $\Sigma^{\rm FOV}_{\rm DM,M31}$ is also more conservative:
less than 85\% of the KZS estimate for this annulus. We further note that for our FOV, CDM models predict
more conservative $\Sigma^{\rm FOV}_{\rm DM,M31}$ values than cored, WDM models. Because WDM
models are less centrally concentrated than CDM models, they are constrained to predict higher densities at larger radii
in order to reach agreement with the total mass of a given halo. For instance, within the radial limits
of our FOV (2.8-6.4 kpc), the dark matter density given by the Burkert profile for M31
\cite{Burkert:1995yz} with the commonly adopted parameter values (as in, e.g., Ref.~\cite{Boyarsky:2007ay})
is $17\% - 29\%$ higher than that of the KZS profile
and $18\% - 41\%$ higher than that of the SBB profile.
Therefore, despite the fact we are constraining a WDM candidate in this paper,
we chose the SBB profile over a WDM profile in the interest of providing a
conservative dark matter mass estimate.
Because of \it{Chandra's}\rm~excellent angular resolution,
we lose less than 5\% of $\Sigma^{\rm FOV}_{\rm DM,M31}$
due to the excision of point sources. However,
to be conservative, we use only $0.95_{~} \Sigma^{\rm FOV}_{\rm DM,M31}$ for $\Sigma^{\rm FOV}_{\rm DM}$
in Eqns.~(\ref{dN_dE_dt_Om}) and (\ref{dN_dE_dt_sin}) to determine
the prospective sterile neutrino signals in the figures below,
and we ignore the contribution from the fraction of the Milky Way dark matter halo within the FOV.
\section{Constraining Sterile Neutrino Decays}
\label{Analysis}
In Fig.~\ref{Galspec}, we show the unresolved emission spectrum from the inner $12'-28'$ of
Andromeda.
To calculate the $\nu_s$ decay signals, we
assume that sterile neutrinos comprise all of the dark matter, i.e.,
$\Omega_{\rm s} = \Omega_{\rm DM} = 0.24$,
and evaluate Eqn.~(\ref{Phixs_nomix}) based on 95\% of the $\Sigma^{\rm FOV}_{\rm DM,M31}$ value
given in Eqn.~(\ref{M_DM}).
We then convert
the sterile neutrino decay fluxes to the same units as those of the
measured spectrum (Counts/sec/keV) as follows:
\begin{eqnarray}
\frac{dN_{\gamma, \rm s}}{dE_{\gamma, \rm s} dt} \left(\Omega_{\rm s}\right) =
\left(\frac{\Phi_{\rm x,s}(\Omega_{\rm s})}{E_{\gamma, \rm s}}\right)
\left(\frac{A_{\rm eff} (E_{\gamma, \rm s})}{\Delta E}\right)
\nonumber \\
= 6.7 \times 10^{-2}~\rm{Counts/sec/keV}
\left(\frac{\rm A_{\rm eff} (E_{\gamma, \rm s})}{100~\rm{cm}^2}\right) \nonumber \\
\times \left(\frac{\Sigma^{\rm FOV}_{\rm DM}}{10^{11}M_\odot \rm Mpc^{-2}}\right)
\left(\frac{\Omega_{\rm s}}{0.24}\right)^{0.813}
\left(\frac{m_s}{\rm~keV}\right)^{1.374},
\label{dN_dE_dt_Om}
\end{eqnarray}
where 1 erg/$E_{\gamma, \rm s} = 1.6 \times 10^{9}(E_{\gamma, \rm s}/\rm{keV})^{-1}$ gives the number
of Counts associated with 1 erg at energy $E_{\gamma, \rm s}$ and
$A_{\rm eff} (E_{\gamma, \rm s})$ and $\Delta E$ are the
effective area\footnote{http://cxc.harvard.edu/proposer/POG/html/ACIS.html} and spectral energy resolution of the
\it{Chandra}\rm~ACIS-I detector, respectively. To realistically simulate detected sterile neutrino decay
``line'' fluxes in Fig.~\ref{Galspec}, we use a Gaussian centered at $E_{\gamma , s} = m_s/2$
with a FWHM of $\Delta E = E_{\gamma , s}/15$, a conservative estimate of the energy
resolution of ACIS-I in imaging mode\footnote{http://heasarc.nasa.gov/docs/cxo/cxo.html}.
\begin{table}[t!]
\caption{The Table shows the distance to Andromeda, the angular range, $\Delta \theta_{\rm FOV}$, of the annular field
of view (FOV) probed by the \it{Chandra}\rm~observations, the dark matter
column density, $\Sigma^{\rm FOV}_{\rm DM}$, enclosed within the FOV (Eqn.~\ref{M_DM}), and the (95\% C.L.)
upper bounds on $m_s$ for Dirac and Majorana sterile neutrinos.}
\begin{ruledtabular}
\begin{tabular}{cc}
Galaxy Name& Andromeda (M31)\\ \hline
Distance (Mpc)& $0.78 \pm 0.02$\\ \hline
$\Delta \theta_{\rm FOV}$ (arcminutes)& $12' - 28'$\\ \hline
$\Sigma^{\rm FOV}_{\rm DM}/10^{11} M_\odot \rm{Mpc}^{-2}$& $0.8 \pm 0.08$\\ \hline
\textbf{$m^{\rm D}_{\rm s}$} (keV) (95\% C.L.)& \textbf{2.4} (Dirac)\\ \hline
\textbf{$m^{\rm M}_{\rm s}$} (keV) (95\% C.L.)& \textbf{2.2} (Majorana)\\ \hline
\end{tabular}
\end{ruledtabular}
\end{table}
\subsection{DW Sterile Neutrino Mass Limits}
\label{mass_limits}
To determine the mass limit imposed by the unresolved emission spectrum of Andromeda, we find the first
bin of energy $E_{\gamma , s} = m_s/2$ for which the sterile neutrino decay signal
(Eqn.~\ref{dN_dE_dt_Om}) at least doubles the amplitude of the measured spectrum in that bin,
$\Delta {\cal F}$:
\begin{equation}
\frac{dN_{\gamma, \rm s}}{dE_{\gamma, \rm s} dt}\left(\Omega_{\rm s}\right) \geq \Delta {\cal
F}.
\label{m_s_limit_criterion}
\end{equation}
The resulting limits: $m^{\rm M}_{\rm s} > 2.2$ keV (Majorana) and $m^{\rm
D}_{\rm s} > 2.4$ keV (Dirac), which are shown in Fig. 2 and Table I, are much more significant than the 95\% C.L.
defined by the ($1_{~}\sigma$) Poisson error bars on the measured points.
(For this data set, Eqn.~(\ref{m_s_limit_criterion}) is equivalent to $\lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle >}\over{\sim}\ $}} 4~\sigma$).
However, because we do not have a
precise understanding of the features of the unresolved emission spectrum, which originate from some
combination of hot gas, unresolved stellar sources, sky and detector backgrounds, and possibly sterile neutrino
decay, we choose a limiting criterion that would remain robust even to 100\% level
fluctuations in the data. (See W06 \cite{Watson:2006}
for a more detailed discussion).
\begin{figure}
\includegraphics[height = .35\textheight, width = .5\textwidth]{M31_CXO_nus_excl_peaks.ps}
\caption {Here we show the \it{Chandra}\rm~unresolved X-ray spectrum emitted from a $12'-28'$ annular
region about the center of the Andromeda galaxy (solid) and the
first statistically significant $\nu_s$ decay
peaks (dashed) at $E_{\gamma, \rm s} = m_{\rm s, lim}/2$ = 1.1 keV (blue) and 1.2 keV (red),
which exclude $m^{\rm M}_{\rm s} > 2.2$ keV (Majorana) and $m^{\rm
D}_{\rm s} > 2.4$ keV (Dirac), respectively (95\% C.L.; Eqn.~\ref{m_s_limit_criterion}).
As a gauge of the statistical significance of our limits, we have included the ($1_{~}\sigma$)
Poisson error bars on the unresolved emission spectrum data points. As a point of comparison to the
potential detection discussed in Ref.~\cite{Loewenstein:2009cm}, we also show
the decay signature that would be produced if the dark matter within the $12'-28'$ region of Andromeda were composed
of 5 keV sterile neutrinos (green, dot-dashed), a possibility that is strongly excluded by the data.
\label{Galspec}}
\end{figure}
\begin{figure}
\includegraphics[height = .35\textheight, width = .5\textwidth]{M31_CXO_nus_fdet_peaks_new.ps}
\caption {Here, to illustrate how difficult it is to distinguish between atomic and anomalous
line features with current detectors, we show the statistical consistency of
the 1.07 keV Ne IX emission peak (\it{Chandra}\rm~data points in black)
and the decay signature of a 2.13 keV Majorana
sterile neutrino (blue, shaded region).
The light blue band shows the 1 $\sigma$ uncertainty range associated
with the \it{Chandra}\rm~data.
\label{detection_features}}
\end{figure}
\subsection{Examining Possible DW Sterile Neutrino Decay Signatures}
\label{detection_tests}
There are a large number of atomic emission lines at energies
just below our exclusion limit, i.e., $E_{\gamma, \rm s} \lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle <}\over{\sim}\ $}} 1$ keV.
Differentiating between these features and anomalous lines will be difficult
if not impossible given the spectral resolution of current detectors.
The diminishing amplitude of sterile neutrino decay signatures at lower
energies, $\frac{dN}{dE dt} \propto E_{\gamma, \rm s}^{1.374}$ (Eqn. \ref{dN_dE_dt_Om})
further exacerbates this problem. Even with a data set such as ours, with a favorably high
ratio of FOV dark matter to astrophysical background, prospective sterile neutrino signals
become comparable to or dwarfed by atomic line features at energies just below 1.1 keV.
To illustrate this point, we examined whether or not any of the atomic emission lines below 1.1 keV in Fig.~2 were
statistically consistent with sterile neutrino decay. We found that this was the case for the Ne IX peak at
1.07 keV. In particular, when we assumed that the Ne IX peak was a sterile neutrino decay line superposed
on a continuum background characterized by the mean flux values of the
bins to the left and right of the peak energy, we found that it was best fit by the decay signature of a 2.13 keV Majorana sterile neutrino (see Fig.~\ref{detection_features}).
This exercise underscores both the spectral resolution and amplitude issues.
If Andromeda's dark matter halo were actually composed of 2.13 keV sterile neutrinos,
they would produce a decay signature at 1.065 keV, but this is clearly indistinguishable
from the 1.07 keV Ne IX feature with the detectors aboard \it{Chandra}.\rm~If the sterile neutrinos
were of even lower mass, the amplitude of their decay signatures would be dwarfed among the
thicket of sub-keV atomic lines, and higher spectral resolution would become indispensible if
we were to have any hope of finding them. One of the few options available in the absence of
such advances is to consider anomalous line ratios, as in Ref.~\cite{Prokhorov:2010us}, but
doing so requires extremely precise knowledge of chemical abundances, plasma temperature,
etc. of the region(s) of the target galaxy being examined. To make progress without such
complications, particularly at $E_{\gamma, \rm s} \lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle <}\over{\sim}\ $}} 1$ keV, a new generation of much
higher spectral resolution detectors is required, as we discuss further in the Sec.~\ref{conclusions}.
\begin{figure}
\includegraphics[height = .4\textheight, width = .5\textwidth]{M31_ms_mix.ps}
\caption {Here we present constraints on $m_s$ as a function of
mixing angle, sin$^2 2\theta$, assuming that all dark matter is comprised of sterile
neutrinos ($\Omega_{\rm s} = 0.24$).
For $L \simeq 10^{-10}$, the thick, solid line corresponds to $\Omega^{\rm DW}_{\rm s}= 0.24\pm 0.04$ in
the Dodelson-Widrow (DW) scenario (Eqn.~\ref{Omega_sin2th}), while the region to the
right corresponds to $\Omega^{\rm DW}_{\rm s} > 0.28$. Three density-production relationships
associated with $\Omega_{\rm s}= 0.3$ and (left to right) $L =$ 0.1, 0.01, and 0.003
are also shown (dotted) \cite{AbaSav}, as is the Shi-Fuller density-production
relationship computed in Ref.~\cite{Laine:2008pg} (dashed).
The three previous radiative decay upper
limits (all 95\% C.L.) are based on \it{Integral}\rm~measurements of the unresolved X-ray
emission from the Milky Way halo \cite{Yuksel:2007xh,Boyarsky:2007ge}, HEAO-1 and
\it{XMM-Newton}\rm~observations of the Cosmic X-ray Background (CXB)\cite{Boyarsky:2005us}, and the most stringent
constraints \cite{Watson:2006} from the many limits imposed by nearby galaxies and clusters
\cite{Aba05b,AbaSav,Boyarsky:2006zi,Watson:2006,Boyarsky:2006fg,Boyarsky:2006ag,Abazajian:2006jc}.
The magenta line shows the recalculated boundary of this exclusion region for
Majorana sterile neutrinos (see text), allowing for direct comparison between our results and those of
Ref. \cite{Watson:2006}.
The most restrictive radiative decay limits, from the present work (also 95\% C.L.), are based on
\it{Chandra}\rm~observations of the Andromeda galaxy.
\label{ms_mix}}
\end{figure}
\subsection{Exclusion Regions in the Mass-Mixing Plane}
\label{exclusion}
To determine the region of the $m_{s~} -$~sin$^{2}2\theta$ (mass-mixing) plane (Fig.~\ref{ms_mix})
that is excluded by the unresolved X-ray spectrum of Andromeda, we convert
Eqn.~(\ref{Phixs}) to Counts/sec/keV:
\begin{eqnarray}
\frac{dN_{\gamma, \rm s}}{dE_{\gamma, \rm s} dt} \left(\sin ^{2}2\theta\right) =
\left(\frac{\Phi_{\rm x,s}(\sin ^{2}2\theta)}{E_{\gamma, \rm s}}\right)
\left(\frac{A_{\rm eff} (E_{\gamma, \rm s})}{\Delta E}\right)
\nonumber \\
= 9.8 \times 10^{-5}~\rm{Counts/sec/keV}
\left(\frac{\rm A_{\rm eff} (E_{\gamma, \rm s})}{100~\rm{cm}^2}\right)\nonumber \\
\times \left(\frac{\Sigma^{\rm FOV}_{\rm DM}}{10^{11}M_\odot \rm Mpc^{-2}}\right)
\left(\frac{\sin ^{2}2\theta}{10^{-10}}\right)
\left(\frac{m_s}{\rm~keV}\right)^{3},
\label{dN_dE_dt_sin}
\end{eqnarray}
and adopt the analog of Eqn.~(\ref{m_s_limit_criterion}) as our exclusion criterion:
\begin{equation}
\frac{dN_{\gamma, \rm s}}{dE_{\gamma, \rm s} dt}\left(\sin ^{2}2\theta \right) \geq \Delta
{\cal F}.
\label{mass_mixing_exclusion}
\end{equation}
Just as we found two mass limits, we also derived two exclusion regions for Dirac and Majorana
sterile neutrinos. The most restrictive region, which was determined by comparing the
\it{Chandra}\rm~unresolved X-ray
spectrum to the Majorana sterile neutrino decay flux, is shown in Fig.~\ref{ms_mix}. The ``indentation'' of our
exclusion region at the highest masses comes about because the effective area of the ACIS-I detector falls
even more steeply than the spectral data at the highest photon energies.
As discussed in the conclusion, a new instrument
with a much larger effective area and superior spectral energy resolution will be required to dramatically improve
upon the radiative decay constraints presented here.
In addition to
our new Andromeda bounds,
(the distinct parts of) three previously determined radiative decay exclusion regions are also shown in
Fig.~\ref{ms_mix}.
The upper exclusion region is based on \it{Integral}\rm~measurements of the unresolved X-ray
emission from the Milky Way halo \cite{Yuksel:2007xh}. We note that these results were
corroborated by a very similar later study \cite{Boyarsky:2007ge}.
The second exclusion region was derived by analyzing
HEAO-1 and \it{XMM-Newton}\rm~observations of the Cosmic X-ray Background (CXB)\cite{Boyarsky:2005us}.
The third region represents the most stringent constraints \cite{Watson:2006} (W06)
from the many limits imposed by nearby galaxies and clusters
\cite{Aba05b,AbaSav,Boyarsky:2006zi,Watson:2006,Boyarsky:2006fg,Boyarsky:2006ag,Abazajian:2006jc}.
We note that if we were to recalculate the exclusion region derived in W06 \cite{Watson:2006} for Majorana rather than Dirac
sterile neutrinos, the boundary of the exclusion region would remain the same (the magenta line shown in Fig. 4).
This is because we would not only need to multiply $\Phi_{\rm x,s}$ by a factor of 2, we would also have
to change the overly optimistic estimate of the spectral energy resolution used in that paper ($\Delta E = E/30$)
to the more realistic value used in this analysis ($\Delta E = E/15$).
Since $\frac{dN_{\gamma, \rm s}}{dE_{\gamma, \rm s} dt} \propto \frac{\Phi_{\rm x,s}}{\Delta E}$ (Eqn. 9),
the factors of 2 offset.
Boyarsky et al. \cite{Boyarsky:2007ay} also conducted a very thorough analysis of an annular
(\it{XMM-Newton}\rm-observed) region of Andromeda ($5' - 13'$).
Unfortunately, because of \it{XMM-Newton's}\rm~poorer
spatial resolution, they were forced to excise almost one fourth of the field of view ($\sim 23$\%) to
remove the emission from point sources, thereby sacrificing potentially signal-producing dark
matter to reduce the astrophysical background. As a result, the unresolved emission spectrum
they generated probed
a much smaller dark matter mass than the \it{Chandra}\rm~unresolved X-ray spectrum we consider here,
and their limits are correspondingly less restrictive ($m_s < 4$ keV).
\section{Conclusions}
\label{conclusions}
In this paper, we used the \it{Chandra}\rm~unresolved X-ray spectrum of the Andromeda galaxy
(M31) to improve the radiative decay constraints on sterile neutrino warm
dark matter.
Assuming either the model described in Refs.~\cite{AFP,AFT,Aba05a,Aba05b} or
Ref.~\cite{Asaka:2006nq}, our analysis requires $m_s < 2.4 \rm\ keV$ (95$\%$ C.L.), in the
least restrictive case (Dirac sterile neutrinos). Because
many recent papers, e.g., \cite{Boyarsky:2007ge,Abazajian:2006jc,Boyarsky:2007ay} have assumed
Majorana sterile neutrinos, we quote the Majorana limit:
\begin{equation}
m_s < 2.2 \rm~keV ~~~(\rm Majorana;~95\% ~\rm C.L.)
\label{M31_12_28_Majorana}
\end{equation}
as our lower bound for the DW scenario, so that comparisons to other studies can be made on equal
footing.
Most of the currently available sterile neutrino mass-mixing
parameter space remains open only for scenarios in which $\Omega_{\rm s} \sim 0.3$ can be
generated at very small mixing. Sterile neutrinos that are, e.g., resonantly produced in the
presence of a large lepton asymmetry ($L \gg 10^{-10}$), e.g., \cite{Shi:1998km,AFP,AbaSav,Laine:2008pg}
or created via Higgs decays \cite{Petraki:2007gq} are still viable. However, the combination of our X-ray
constraints with those from nearby sources
are also able to partially restrict the $L=0.003$ resonant production
model \cite{AbaSav} and the Shi-Fuller model \cite{Laine:2008pg} for sterile neutrino masses in the
10 to 24 keV range (Fig. 4), and the \it{Integral}\rm~measurements of the Milky Way halo
\cite{Yuksel:2007xh,Boyarsky:2007ge} exclude
all the alternative production scenarios shown in Fig.~\ref{ms_mix} above $m_s = 40$ keV.
Even without cosmological small scale structure bounds, the DW scenario remains
viable only between the Tremaine-Gunn bound \cite{TG:1979} and our limit from this work,
0.4 keV $< m_s <$ 2.2 keV, which interestingly falls within the range of dark matter particle
masses that best explains the core of the Fornax Dwarf Spheroidal galaxy \cite{Strigari:2006ue}.
This result underscores the need to continue to carefully and independently pursue all possible constraints
on sterile neutrino properties. However, as noted above,
because of the decreasing sterile neutrino signal at lower $m_s$ values
and the large number of atomic emission features at energies $\lower0.6ex\vbox{\hbox{$ \buildrel{\textstyle <}\over{\sim}\ $}} 1$ keV,
it will be difficult to improve upon the 2.2 keV limit we have presented here with existing
X-ray detectors.
One of the most promising routes toward improved radiative constraints
includes the use
of much higher spectral resolution instruments than those currently available, as discussed, e.g.,
in Refs.~\cite{Abazajian:2006jc,Boyarsky:2006hr,Abazajian:2009hx,Herder:2009im}.
The International X-ray Observatory
(\it{IXO}\footnote{\rm http://ixo.gsfc.nasa.gov}\rm ),~for instance, which is scheduled for launch in 2021,
will have a FOV comparable to \it{Chandra}\rm~detectors ($\sim$~18' for photon energies ranging from 0.1-15 keV),
but $\sim$~100 times their effective area, $\sim$~10 times their spectral resolution, and $\sim$~10 times lower
instrumental background. Based on these specifications, a megasecond observation of Andromeda with IXO will have
the capacity to fully test the 4 alternative sterile neutrino production scenarios shown in Fig.~\ref{ms_mix}
over the entire range of mass-mixing parameter space for which they remain viable \cite{Abazajian:2009hx}.
\acknowledgments
We thank M. Garcia, S. Murray, and Q.D. Wang, the principal investigators responsible for conducting
the Chandra observations of M31 used in this work. We also thank John Beacom and Hasan Y{\"u}ksel
for extensive and very helpful discussions.
CRW and NP acknowledge support from the Millikin University
Department of Physics and Astronomy. ZL acknowledges support from SAO grant GO0-11098B.
|
\section{Introduction}
The standard definition of the thermodynamic equilibrium quantities in a canonical
ensemble relies on the assumption that the coupling between the system of
interest and its thermal environment is arbitrarily weak. In the real world,
this approximation never strictly holds and important deviations from the usual
picture can be expected in particular for systems on the nanoscale. Beyond
the study of effects of finite coupling to the environment on the thermodynamic
properties of system, the more fundamental question arises of how to properly
define quantities like an internal energy, a specific heat or an entropy in
such a situation \cite{hangg06}. The last five years or so have seen considerable
activities addressing this issue and related ones
\cite{ford07,hangg08,hoerh08,wang08,ingol09,campi09a,campi09,gelin09,%
bandy09,kumar09,campi10,datta10,bandy10,bandy10a,kim10,campi11,willi11,haseg11b}.
In the quantum regime, the thermodynamic equilibrium properties not only depend
on the strength of the coupling to the environment. Even more importantly, it
turns out that the very definition of these quantities is no longer unique
\cite{hangg06}. As an example, we consider here the specific heat which for a
damped quantum system can be defined in at least two different ways.
The first approach defines the internal energy of the system as expectation
value of the system Hamiltonian taken with respect to the thermal ensemble of
the coupled system. By the latter we mean here and in the following the coupled
complex of system (S) and heat bath (B). Quantities related to the coupled
system will be indicated by a subscript ``S+B''. Based on the definition of the
internal energy just introduced, one obtains the associated specific heat by
taking the derivative with respect to temperature. Obviously, in the absence
of a coupling between system and bath, the specific heat defined in this way
reduces to the standard specific heat of the system.
An alternative approach is based on the reduced partition function of the
system defined as \cite{dittr98,weiss99,ingol02}
\begin{equation}
\label{eq:reducedPartitionFunction}
\mathcal{Z} = \frac{\mathcal{Z}_\mathrm{S+B}}{\mathcal{Z}_\mathrm{B}}\,.
\end{equation}
In the absence of a coupling between the system and its environment, $\mathcal{Z}$
equals the partition function of the system. Even though we thus can
consider $\mathcal{Z}$ to be a quantity associated with the
system, we omit a subscript ``S'' in order to emphasize the fact that
$\mathcal{Z}$ also takes into account the coupling between system and bath.
Here, we will employ the ratio (\ref{eq:reducedPartitionFunction}) only in
the context of equilibrium thermodynamics, but recently it has also been used in the
discussion of fluctuation theorems for open quantum systems in nonequilibrium
situations \cite{campi09a,campi09,campi11}.
The partition function (\ref{eq:reducedPartitionFunction}) can be used to
define thermodynamic equilibrium quantities by means of the standard relations
valid in the absence of a coupling between the system and its environment. We
thus obtain a second expression for the specific heat reading
\begin{equation}
\label{eq:csfromzred}
C = k_\mathrm{B}\beta^2
\frac{\partial^2\ln(\mathcal{Z})}{\partial\beta^2}\,,
\end{equation}
where $\beta=1/k_\mathrm{B}T$. In the absence of a coupling between system and
bath, this definition of a specific heat also reduces to the usual definition
and thus agrees with our first definition in the limit of vanishing coupling.
Again we omit a subscript ``S'' to underline the dependence of the specific
heat on the coupling, but we do so also in view of the interpretation
(\ref{eq:diffcs}) given below which emphasizes the role of the heat bath.
It should be noted that, in general, the two specific heats just defined differ
for finite coupling to the environment already in the leading high-temperature
corrections to the classical specific heat. For a bilinear coupling between
system and bath, the classical specific heats still agree
\cite{hangg06,hangg08} while this need no longer be the case for anharmonic
couplings \cite{gelin09}.
The difference between the two approaches is most spectacular
in situations where the specific heat (\ref{eq:csfromzred}) becomes negative in
contrast to the positive specific heat based on the expectation value of the
system Hamiltonian \cite{hangg06,hangg08}. In the following, we exclusively consider
the specific heat obtained from the reduced partition function so that
the symbol $C$ will always refer to (\ref{eq:csfromzred}).
Although the reduced partition function (\ref{eq:reducedPartitionFunction})
appears to be a rather formal starting point, the quantities derived from it
have a clear physical meaning. Due to the fact that thermodynamic equilibrium
properties depend on the logarithm of the partition function, these quantities
in the presence of a finite coupling to a heat bath possess a natural
interpretation in terms of the difference between the quantity evaluated for
the coupled system and the quantity evaluated for the bath alone
\cite{ingol09,ford85}. For the specific heat (\ref{eq:csfromzred}), we thus
have
\begin{equation}
\label{eq:diffcs}
C = C_\mathrm{S+B}-C_\mathrm{B}\,.
\end{equation}
The two quantities involved in this difference, $C_\mathrm{S+B}$ and
$C_\mathrm{B}$, constitute measurable quantities. Even though one may invoke a
superbath imposing the temperature on the bath and on the coupled system, the
coupling to the superbath may be assumed to be negligible. Therefore, we are
certainly allowed to employ the standard thermodynamic definition for the
quantities on the right-hand side of (\ref{eq:diffcs}). $C$ thus is not a
specific heat in the proper sense but it is the natural replacement for a
specific heat in a situation where the system of interest is coupled with
non-negligible strength to an environment. The approach discussed here was
recently employed in the analysis of thermal data obtained for a metal
phosphate compound \cite{swain11}. For the sake of simplicity, we will
occasionally refer to $C$ as specific heat of the system. In doing so, we
should however keep in mind the preceding discussion.
With the above interpretation of the specific heat (\ref{eq:csfromzred}),
negative values should not give rise to concerns with respect to thermodynamic
instabilities. In fact, there is no reason why the difference of two positive
quantities should be positive. The situation considered here should therefore
be distinguished from the negative specific heat appearing e.g. in self-gravitating
systems within a microcanonical description \cite{lynde68,thirr70,thirr03}.
Apart from the free damped quantum particle which we focus on in this paper,
negative entropies and/or specific heats have been discussed e.g. in the
context of the Casimir effect
\cite{klimc06,milto08,hoye07,ingol09a,intra09,canag10a,borda10a,rodri11}, Kondo
systems \cite{flore04,zitko08}, XY spin chains \cite{campi10}, two-level
fluctuators \cite{campi09} and energy transport in proteins \cite{sulai10}.
While for a single harmonic oscillator coupled to an Ohmic bath
\cite{hangg06,hangg08} or to a single bath oscillator \cite{ingol09} the
specific heat was found to be positive, this is not necessarily the case if a
system consisting of several harmonic oscillators is coupled to a finite
environment \cite{haseg11b}.
In the treatment of dissipative systems, the system degree of freedom usually
is in the focus of interest. On the other hand, the difference
(\ref{eq:diffcs}) of specific heats, even though conceptually associated with
the system degree of freedom, provides the motivation to take a different point
of view, namely the one of the environment. The question answered by the
specific heat (\ref{eq:diffcs}) really is: How does the specific heat of the
environment change when a system degree of freedom is coupled to it?
In the following we shall address this question by studying the change in the
spectral density of the environment when a system degree of freedom is coupled
to it. In Section~\ref{sec:freeParticleHarmonicOscillators} we briefly
introduce the model for the damped free quantum particle which will allow us to
obtain the change in the spectral density of the environment in
Section~\ref{sec:changeSpectralDensity}. From these results, we derive in
Section~\ref{sec:thermodynamicProperties} the thermodynamic properties of the
system in the sense explained above. We will show that these results can indeed
be expressed in terms of properties of the system degree of freedom. In
Section~\ref{sec:missingMass} we will give a physical interpretation of the
condition under which the specific heat (\ref{eq:diffcs}) of the free damped
quantum particle becomes negative. Finally, we present our conclusions in
Section~\ref{sec:conclusions}.
\section{Free particle coupled to harmonic oscillators}
\label{sec:freeParticleHarmonicOscillators}
As our model for the study of the appearance of a negative specific heat we consider
a particle of mass $M$ which is bilinearly coupled to a set of harmonic oscillators
constituting the environment. The corresponding Hamiltonian is given by
\begin{equation}
\label{eq:hamiltonian}
H = \frac{P^2}{2M} + \sum_{n=1}^\infty\left[\frac{p_n^2}{2m_n}+\frac{m_n\omega_n^2}{2}
(Q-x_n)^2\right]\,.
\end{equation}
The sum describes the environmental oscillators of mass $m_n$ and frequency $\omega_n$,
the bilinear coupling and a potential renormalization. The latter ensures that the
effective equation of motion of the system position $Q$ corresponds indeed to
the Langevin equation of a free damped particle taking the form
\begin{equation}
\label{eq:LangevinEquation}
M\ddot Q+M\int^t\mathrm{d}s\gamma(t-s)\dot Q(s) = \xi(t)\,.
\end{equation}
Below, we will relate the damping kernel $\gamma(t)$ to the properties of the
environment. The properties of the noise term $\xi(t)$ are irrelevant for our
discussion so that we refer the reader to the literature for details
\cite{dittr98,weiss99,ingol02,hangg05}. The coupling constant of the bilinear
term in $x_n$ and $Q$ in (\ref{eq:hamiltonian}) has been expressed in terms of
the masses and frequencies of the environmental oscillators which is possible
without loss of generality \cite{hakim85,grabe88}. The Hamiltonian is manifestly
invariant under spatial translations of all degrees of freedom, confirming once
more that we are treating a free damped particle.
In our discussion, two spectral densities play a key role and it is important
to carefully distinguish them. We refer to the first quantity as spectral
density of eigenmodes $\rho$. This quantity is defined in terms of the
eigenfrequencies of the quadratic Hamiltonian (\ref{eq:hamiltonian}). For the
bath alone, we have the spectral density of eigenmodes
\begin{equation}
\label{eq:spectralDensityOfEigenmodes0}
\rho_\mathrm{B}(\omega) = \sum_{n=1}^\infty\delta(\omega-\omega_n)\,.
\end{equation}
Correspondingly, we define a spectral density of eigenmodes
$\rho_\mathrm{S+B}(\omega)$ in which the frequencies $\omega_n$ appearing in
(\ref{eq:spectralDensityOfEigenmodes0}) are replaced by the eigenfrequencies of
the coupled system described by (\ref{eq:hamiltonian}). The difference
$\rho_\mathrm{S+B}-\rho_{B}$ will form the basis for the derivation of various
thermodynamic quantities in Section~\ref{sec:thermodynamicProperties}.
In addition, in order to characterize the properties of the heat bath and its
coupling to the system, a quantity commonly referred to as spectral density
of bath oscillators \cite{calde83}
\begin{equation}
\label{eq:spectralDensityOfBathOscillators}
J(\omega) = \frac{\pi}{2}\sum_{n=1}^\infty m_n\omega_n^3\delta(\omega-\omega_n)
\end{equation}
is introduced. In contrast to the spectral density of eigenmodes
(\ref{eq:spectralDensityOfEigenmodes0}) in the absence of coupling, the spectral
density of bath oscillators (\ref{eq:spectralDensityOfBathOscillators}) not
only depends on the frequencies of the bath oscillators but also on their
masses. For our choice of coupling constants (cf. discussion below
(\ref{eq:LangevinEquation})) this is tantamount to saying that
(\ref{eq:spectralDensityOfBathOscillators}) depends also on the coupling
constants between system degree of freedom and bath oscillators. $J(\omega)$
contains all information about the heat bath required to describe the
properties of the damped quantum system.
Although it is not obvious, it will turn out that even within our bath-centered
approach all quantities of interest will eventually be expressible in terms of
the spectral density of bath oscillators
(\ref{eq:spectralDensityOfBathOscillators}). This already indicates that even
though we take the point of view of the environment, we will learn something
about the properties of the system.
In our results, it will be natural to express the dependence on $J(\omega)$ in
terms of the Laplace transform of the damping kernel $\gamma(t)$ appearing in
(\ref{eq:LangevinEquation}). The two quantities are related by means of
\cite{dittr98,weiss99,ingol02}
\begin{equation}
\label{eq:gammaHat}
\hat\gamma(z) = \frac{2}{\pi M}\int_0^\infty\mathrm{d}\omega\frac{J(\omega)}{\omega}
\frac{z}{\omega^2+z^2}\,.
\end{equation}
For the purpose of our discussion, we will restrict ourselves to so-called Ohmic
damping which implies that the spectral density of bath oscillators is continuous
and linear in the frequency at least for small frequencies. At high frequencies,
the spectral density of bath oscillators will typically be suppressed with respect
to this linear behavior. To be specific, in cases where such a cutoff
is of relevance, we will employ the so-called Drude model with
\begin{equation}
\label{eq:densityDrude}
J_\mathrm{D}(\omega) = M\gamma\omega
\frac{\omega_\mathrm{D}^2}{\omega^2+\omega_\mathrm{D}^2}\,.
\end{equation}
Here, the low-frequency behavior is characterized by the damping constant $\gamma$
and $\omega_\mathrm{D}$ is the cutoff frequency. The corresponding Laplace transform
of the damping kernel reads
\begin{equation}
\label{eq:gammaHatDrude}
\hat\gamma_\mathrm{D}(z) = \frac{\gamma\omega_\mathrm{D}}{z+\omega_\mathrm{D}}\,.
\end{equation}
Because the damping constant $\gamma$ combines with the inverse temperature
$\beta$ to form a dimensionless quantity $\hbar\beta\gamma$ and thus
merely sets the temperature scale, the cutoff frequency $\omega_\mathrm{D}$ will be
an important parameter determining the properties of the damped free particle.
\section{Change in the spectral density of bath oscillators}
\label{sec:changeSpectralDensity}
In our environment-centered approach, we start out with a set of uncoupled environmental
oscillators with frequencies $\omega_n$. In addition, we have one system degree of freedom
corresponding to an undamped free particle. We now couple the environmental oscillators
to the free particle as prescribed by the Hamiltonian (\ref{eq:hamiltonian}). As discussed
in the previous section, the system and environment together are translationally invariant.
In the spirit of Ullersma's analysis of the damped harmonic oscillator
\cite{uller66} we determine the eigenmode spectrum of the free particle coupled
to its environment. Diagonalizing the Hamiltonian, we recover a zero-mode
which replaces the zero-mode corresponding to the uncoupled free particle. We
thus concentrate on the non-zero frequencies which due to the coupling are
shifted with respect to the original environmental oscillator frequencies. The
new frequencies are obtained as solutions $\Omega$ of the equation
\begin{equation}
\label{eq:evCondition}
\sum_{n=1}^\infty \frac{m_n\omega_n^2}{\Omega^2-\omega_n^2} = M\,.
\end{equation}
For the following intermediate steps it is convenient to explicitly consider an
equidistant set of discrete frequencies $\omega_n=n\Delta$ with $n=1, 2, \dots$
for the environmental oscillators. Later, we will take the limit of vanishing frequency
spacing $\Delta$ to recover a continuous spectral density of environmental oscillators.
In view of (\ref{eq:spectralDensityOfBathOscillators}), we have the relation
\begin{equation}
\int_0^\infty\mathrm{d}\omega\,J(\omega)f(\omega) = \frac{\pi}{2}\sum_{n=1}^\infty
m_n\omega_n^3f(n\Delta)
\end{equation}
which holds for any function $f$ for which the integral and the sum exist. If, on the
other hand, we replace the integral on the left-hand side by a Riemann sum with step
width $\Delta$, we can express the masses $m_n$ of the environmental oscillators in terms
of the continuous spectral density of bath oscillators $J(\omega)$ according to
\begin{equation}
m_n = \frac{2}{\pi}\frac{J(n\Delta)}{(n\Delta)^3}\Delta\,.
\end{equation}
The eigenfrequency condition (\ref{eq:evCondition}) thus becomes
\begin{equation}
\sum_{n=1}^\infty\frac{J(n\Delta)}{n[\Omega^2-(n\Delta)^2]} = \frac{\pi}{2}M\,.
\end{equation}
In order to treat frequencies $\Omega$ close to an unperturbed environmental frequency
$n\Delta$ correctly, we rewrite this condition as
\begin{equation}
\begin{aligned}
&\Delta\frac{J(\Omega)}{\Omega}\sum_{n=1}^\infty\frac{1}{\Omega^2-(n\Delta)^2}\\
&\quad+\Delta\sum_{n=1}^\infty\frac{1}{\Omega^2-(n\Delta)^2}\left(\frac{J(n\Delta)}{n\Delta}
-\frac{J(\Omega)}{\Omega}\right)=\frac{\pi}{2}M
\end{aligned}
\end{equation}
Performing the first sum exactly and replacing the second sum by an integral, we finally
obtain
\begin{equation}
\label{eq:newEigenvaluesCondition}
\cot\left(\frac{\pi\Omega}{\Delta}\right)-\frac{\Delta}{\pi\Omega} = g(\Omega)
\end{equation}
with
\begin{equation}
\label{eq:functiong}
g(\Omega) = \frac{M\Omega}{J(\Omega)}\big(\Omega+\mathrm{Im}\hat\gamma(\mathrm{i}\Omega)\big)\,.
\end{equation}
The coupling of the environment to the system degree of freedom modifies the
original spacing $\Delta$ between adjacent environmental eigenfrequencies
yielding a new spacing $\Delta+\epsilon$. The correction $\epsilon$ depends on
the spectral bath density via the function $g$ defined in (\ref{eq:functiong})
and can be shown to be of order $\Delta^2$. Exploiting the latter fact and
making use of the addition theorem of the cotangent, one determines $\epsilon$
from (\ref{eq:newEigenvaluesCondition}). For the change in the spectral density
of eigenmodes, one thus finds as a central result
\begin{equation}
\label{eq:rhoMinusRho0}
\rho_\mathrm{S+B}(\Omega)-\rho_\mathrm{B}(\Omega) =
\frac{1}{\Delta+\epsilon(\Omega)}-\frac{1}{\Delta}
= \frac{1}{\pi}\frac{g'(\Omega)}{1+g(\Omega)^2}\,.
\end{equation}
Here, the prime denotes a derivative with respect to the argument.
Further insight into the change of the spectral density of eigenmodes can be
obtained by specifying the properties of the bath. We choose a Drude model
with the spectral density of environmental oscillators given by
(\ref{eq:densityDrude}) and a Laplace transform of the damping kernel as specified
in (\ref{eq:gammaHatDrude}). From (\ref{eq:rhoMinusRho0}) one obtains
\begin{equation}
\label{eq:lorentziansReal}
\rho_\mathrm{S+B}-\rho_\mathrm{B} = \frac{1}{\pi}\left[\frac{\omega_1}{\Omega^2+\omega_1^2}
+\frac{\omega_2}{\Omega^2+\omega_2^2}
-\frac{\omega_\mathrm{D}}{\Omega^2+\omega_\mathrm{D}^2}\right]
\end{equation}
where
\begin{equation}
\label{eq:omega12}
\omega_{1,2} = \omega'\pm\mathrm{i}\omega''
= \frac{\omega_\mathrm{D}}{2}\left(1\pm\sqrt{1-\frac{4\gamma}{\omega_\mathrm{D}}}\right)
\end{equation}
and $\omega_\mathrm{D}$ is the Drude frequency providing the high-fre\-quency cutoff.
These three frequencies are the eigenfrequencies associated with the deterministic
version of the Langevin equation (\ref{eq:LangevinEquation}) for a Drude damping
kernel \cite{hangg08}.
The change in the spectral density of eigenmodes is shown as thick solid line in
Fig.~\ref{fig:rho_lorentz} for various values of $\omega_\mathrm{D}/\gamma$.
The dotted lines represent the Lorentzian contributions according to
(\ref{eq:lorentziansReal}). For the sake of clarity, they are depicted also in
the grey regions of negative frequencies which are irrelevant for the spectral
density of eigenmodes. As long as the cutoff frequency is sufficiently high,
$\omega_\mathrm{D}>4\gamma$, (\ref{eq:lorentziansReal}) describes the sum of
three Lorentzians centered at zero frequency. An example for this situation is
given in Fig.~\ref{fig:rho_lorentz}a. Coupling the system degree of freedom to the
environment then leads to an increase of the spectral density at zero frequency
by $(\omega_\mathrm{D}-\gamma)/\pi\gamma\omega_\mathrm{D}$. For smaller cutoff
frequencies, if $\omega_\mathrm{D}<4\gamma$ like in the cases depicted in
Figs.~\ref{fig:rho_lorentz}b and c, the two frequencies (\ref{eq:omega12}) become
complex and the first two Lorentzians in (\ref{eq:lorentziansReal}) are now
centered at nonzero frequencies. As a consequence, the suppressing effect of
the third Lorentzian on the spectral density of eigenmodes becomes relevant.
For $\omega_\mathrm{D}<\gamma$ one finds in fact a suppression of the spectral
density at zero frequency as is shown in Fig.~\ref{fig:rho_lorentz}c for
$\omega_\mathrm{D}/\gamma=0.1$. It is this suppression which leads to the
negative specific heat as we will discuss in the following section.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\columnwidth]{rho_lorentz}
\end{center}
\caption{The change in the spectral density of eigenmodes
$\rho_\mathrm{S+B}-\rho_\mathrm{B}$ induced by the coupling to the system degree
of freedom is depicted as thick solid line for Drude damping with (a)
$\omega_\mathrm{D}/\gamma=5$, (b) $\omega_\mathrm{D}/\gamma=1$,
and (c) $\omega_\mathrm{D}/\gamma=0.1$. The contributions of the three Lorentzians
according to (\ref{eq:lorentziansReal}) are represented as dotted lines. Only
the white regions of positive frequencies are of relevance for the spectral
density of eigenmodes.}
\label{fig:rho_lorentz}
\end{figure}
Before doing so, we would like to emphasize that although the change in the
spectral density of eigenmodes obtained in this section is mainly related to
properties of the environment, the results are determined by properties of the
damped system. In the general result (\ref{eq:rhoMinusRho0}) the environment
appears only through the spectral density of bath oscillators
(\ref{eq:spectralDensityOfBathOscillators}) which is sufficient to provide a
reduced description of the damped system. In the specific case of Drude
damping, the three eigenfrequencies of the damped system turn out to be
sufficient for the complete description of the change in the spectral density
of eigenfrequencies.
Despite being closely related to the properties of the damped system, the
change of the spectral density of eigenmodes discussed here should not be
confused with the density of states which can be obtained from the reduced
partition function (\ref{eq:reducedPartitionFunction}) by its inverse Laplace
transform \cite{hanke95}. Because the density of states defined in such a way
is linearly related to the reduced partition function, it cannot be interpreted
as a difference of a property of the coupled system on the one hand and the
heat bath on the other hand \cite{hangg08}.
\section{Thermodynamic properties of the system}
\label{sec:thermodynamicProperties}
The change of the specific heat (\ref{eq:csfromzred}) due to the coupling of
the system degree of freedom to the environment can be obtained from the change
(\ref{eq:rhoMinusRho0}) of the spectral density of eigenfrequencies as
\begin{equation}
\label{eq:csGeneral}
C = \int_0^\infty\mathrm{d}\omega
\big(\rho_\mathrm{S+B}(\omega)-\rho_\mathrm{B}(\omega)\big)C_\mathrm{ho}(\omega)
\end{equation}
where
\begin{equation}
\label{€q:cho}
C_\mathrm{ho}(\omega) = k_\mathrm{B}
\left(\frac{\hbar\beta\omega}{2\sinh(\hbar\beta\omega/2)}\right)^2\,.
\end{equation}
is the specific heat of a harmonic oscillator with frequency $\omega$.
Eq.~\ref{eq:csGeneral} implies a negative specific heat $C$ at low temperatures
if the spectral density of eigenmodes is suppressed for small frequencies due
to the coupling of the system degree of freedom to the bath.
Assuming Ohmic damping, i.\,e. $J(\omega)\sim\omega$ at low frequencies,
we can obtain the specific heat at low temperatures by inserting the change
in the spectral density of eigenmodes (\ref{eq:rhoMinusRho0}) into
(\ref{eq:csGeneral}). Together with (\ref{eq:functiong}) one finds
\begin{equation}
\label{eq:csLowTemperature}
\frac{C}{k_\mathrm{B}} = \frac{\pi}{3}\frac{1+\hat\gamma^\prime(0)}
{\hat\gamma(0)}\frac{k_\mathrm{B}T}{\hbar} + \mathcal{O}(T^3)
\end{equation}
in agreement with the findings of Ref.~\cite{hangg08}. For
\begin{equation}
\label{eq:negativityCondition}
\hat\gamma^\prime(0) <-1\,,
\end{equation}
we thus obtain a negative specific heat at low temperatures. A physical
interpretation of this result will be given in Section~\ref{sec:missingMass} below.
More explicit results can be derived for the case of Drude damping. There, the
integral in (\ref{eq:csGeneral}) can be evaluated analytically and one finds
\begin{equation}
\begin{aligned}
\label{eq:csDrude}
\frac{C}{k_\mathrm{B}} &=
\left(\frac{\hbar\beta\omega_1}{2\pi}\right)^2\psi^\prime
\left(\frac{\hbar\beta\omega_1}{2\pi}\right)
+\left(\frac{\hbar\beta\omega_2}{2\pi}\right)^2\psi^\prime
\left(\frac{\hbar\beta\omega_2}{2\pi}\right)\\
&\qquad-\left(\frac{\hbar\beta\omega_\mathrm{D}}{2\pi}\right)^2\psi^\prime
\left(\frac{\hbar\beta\omega_\mathrm{D}}{2\pi}\right)
-\frac{1}{2}\,,
\end{aligned}
\end{equation}
where $\psi^\prime(z)$ denotes the digamma function. This result is in
agreement with the expression obtained by proceeding according to the point of
view of the system \cite{hangg08} and starting with the reduced partition function
(\ref{eq:reducedPartitionFunction}). The two approaches therefore are equivalent,
but the approach presented here gives additional insight through the change
in the spectral density of eigenvalues (\ref{eq:rhoMinusRho0}).
Instead of obtaining the internal energy and the specific heat from the reduced
partition function, we can use our result (\ref{eq:csDrude}) for the specific
heat to obtain the other two quantities. In the limit of vanishing coupling to
the heat bath, the internal energy $U$ is related to the specific heat $C$ by means of
\begin{equation}
C=-\beta^2\frac{\partial U}{\partial\beta}\,.
\end{equation}
In view of (\ref{eq:diffcs}) we thus obtain from (\ref{eq:csDrude}) by means of an
integration the difference of internal energies induced by the system-bath coupling
\begin{equation}
\begin{aligned}
U &= U_\mathrm{S+B}-U_\mathrm{B}\\
&= -\frac{\hbar\omega_1}{2\pi}\psi\left(\frac{\hbar\beta\omega_1}{2\pi}\right)
-\frac{\hbar\omega_2}{2\pi}\psi\left(\frac{\hbar\beta\omega_2}{2\pi}\right)\\
&\qquad+\frac{\hbar\omega_\mathrm{D}}{2\pi}\psi\left(\frac{\hbar\beta\omega_\mathrm{D}}{2\pi}\right)
-\frac{1}{2\beta}\,.
\end{aligned}
\end{equation}
This result is only determined up to a constant of integration which, by
comparison with the known result \cite{hangg08}, turns out to vanish. In
particular, at zero temperature, we thus find
\begin{equation}
U_0 = \frac{\hbar\omega_1}{2\pi}\ln\left(\frac{\omega_\mathrm{D}}{\omega_1}\right)
+\frac{\hbar\omega_2}{2\pi}\ln\left(\frac{\omega_\mathrm{D}}{\omega_2}\right)\,.
\end{equation}
Furthermore, by means of the relation between the internal energy $U$ and the partition
function $\mathcal{Z}$
\begin{equation}
U = -\frac{\mathrm{d}}{\mathrm{d}\beta}\ln(\mathcal{Z})
\end{equation}
one reproduces the correct temperature dependence of the ratio of the partition functions
of system and bath on the one hand and bath alone on the other hand \cite{hangg08}
\begin{equation}
\mathcal{Z}\sim\frac{1}{\beta^{1/2}}
\frac{\Gamma\left(1+\dfrac{\hbar\beta\omega_1}{2\pi}\right)
\Gamma\left(1+\dfrac{\hbar\beta\omega_2}{2\pi}\right)}
{\Gamma\left(1+\dfrac{\hbar\beta\omega_\mathrm{D}}{2\pi}\right)}\,.
\end{equation}
This result leaves a prefactor undetermined which for the thermodynamical
equilibrium quantities is irrelevant. The temperature dependence, however,
agrees with the result obtained by means of other techniques, e.g. the
path integral approach.
\section{The missing mass of bath oscillators}
\label{sec:missingMass}
In the previous section, we have found that a negative specific heat occurs
provided the condition (\ref{eq:negativityCondition}) is satisfied.
This rather formal condition can be given a physical meaning. In order to
avoid an infrared divergence in the limit of vanishing argument $z$, we first
express the Laplace transform of the damping kernel (\ref{eq:gammaHat}) as
\begin{equation}
\label{eq:gammaHatDiff}
\hat\gamma(z) = \hat\gamma(0)+\frac{2}{\pi}\int_0^\infty\mathrm{d}\omega
\left(\frac{J(\omega)}{M\omega}-\hat\gamma(0)\right)\frac{z}{\omega^2+z^2}\,.
\end{equation}
With (\ref{eq:negativityCondition}) and (\ref{eq:gammaHatDiff}) the condition
for a negative specific heat then becomes
\begin{equation}
\label{eq:negativityCondition2}
\frac{2}{\pi}\int_0^\infty\mathrm{d}\omega\frac{M\hat\gamma(0)\omega-J(\omega)}
{\omega^3}>M\,.
\end{equation}
Observing that according to (\ref{eq:spectralDensityOfBathOscillators}) the total
mass of the bath oscillators can be obtained from the spectral density of bath
oscillators $J(\omega)$ as \cite{hakim85,grabe88}
\begin{equation}
\mathcal{M} = \sum_{n=1}^\infty m_n = \frac{2}{\pi}\int_0^\infty\mathrm{d}\omega
\frac{J(\omega)}{\omega^3}
\end{equation}
the condition (\ref{eq:negativityCondition2}) can be expressed as
\begin{equation}
\label{eq:massCondition}
\Delta\mathcal{M}_\mathrm{B}>M\,.
\end{equation}
Here, $\Delta\mathcal{M}_\mathrm{B}$ is defined by the left-hand side of
(\ref{eq:negativityCondition2}) and refers to the total mass of oscillators
which are missing in the actual bath with respect to a strictly Ohmic reference
bath where $J(\omega)=M\hat\gamma(0)\omega$ for all frequencies. We note that
while due to the infrared divergence mentioned above the total mass is infinite
for both baths, the difference $\Delta\mathcal{M}_\mathrm{B}$ is finite.
We thus arrive at the following interpretation. Coupling a system degree of
freedom of zero frequency to the bath oscillators tends to shift the
frequencies of the latter to larger values. However, for a strictly Ohmic
environment, i.e. in the absence of a finite cutoff in the spectral density
of bath oscillators, the ensemble of bath oscillators resists a reduction of
the spectral density of eigenmodes at low frequencies. A suppression only becomes
possible if one eliminates a number of bath oscillator with a total mass at
least as large as the mass associated with the system degree of freedom. The
bath then is no longer able to resist the ``pressure'' of the system degree of
freedom pushing the eigenmodes to higher frequencies.
\section{Conclusions}
\label{sec:conclusions}
In order to achieve a better understanding of the unusual thermodynamic
properties of a free Brownian quantum particle, we have employed a somewhat
uncommon approach to the analysis of dissipative quantum systems by taking the
perspective of the environment. This approach is, however, quite natural if
thermodynamic quantities are defined in terms of a reduced partition function
of the system because they actually refer to the change of these quantities when
the system degree of freedom is coupled to the heat bath.
We have analyzed the modification of the bath spectrum induced by the coupling
to the system degree of freedom. Interestingly, the spectral density of bath
oscillators is sufficient to describe the change in the bath spectrum. As a
consequence the latter depends only on properties of the damped system as has
been exemplified by means of a Drude-type damping. Starting from the change in
the bath spectrum, expressions for the specific heat, the internal energy and
the reduced partition function obtained previously from a system-based approach
have been reproduced.
The low-temperature behavior of the specific heat of the free damped particle
is determined by the shift of the low-frequency environmental oscillators
induced by the coupling to the system degree of freedom. If the ratio of cutoff
frequency and damping strength is sufficiently small, the system degree of
freedom succeeds in suppressing the spectral density of eigenmodes at low
frequencies by shifting the bath modes to higher frequencies. As a result, at
low temperatures the specific heat of the heat bath is lowered if the system
degree of freedom is attached. In contrast, for larger cutoff frequencies, the
high-frequency oscillators of the environment act against the tendency of
the system to shift the bath modes to higher frequencies. Then, the specific
heat remains positive for all frequencies.
The condition to be satisfied by the environment to allow for a negative
specific heat at low temperatures has been shown to have a physical
interpretation. The anomaly in the specific heat appears if the mass missing
in the environment with respect to a strictly Ohmic reference bath exceeds the
mass associated with the system degree of freedom. Then, the bath oscillators
are not strong enough to resist the zero-frequency degree of freedom which,
when coupled to the environment, tends to increase their frequencies. Otherwise
the spectrum of bath modes is sufficiently stiff to prevent the spectral
density of eigenmodes from being reduced.
\begin{acknowledgement}
The author is grateful to Michele Campisi, Peter H\"anggi, Astrid Lambrecht,
Serge Reynaud, Peter Talkner, and Juan Diego Urbina for useful discussions.
This work has been supported by the DAAD through the PROCOPE program.
\end{acknowledgement}
|
\section{Numerical Methods}
We have developed a hybrid code to simultaneously
calculate the evolution of an accreting first generation
(proto-)star and the flow of the surrounding
primordial material.
We solve the accretion flow with a radiation hydrodynamic
code, whereby the central (proto-)star is represented as a
sink cell. This sink cell grows in mass by inflow through its
boundary and it influences the surrounding material flow via
gravity and by emitting radiation.
We follow the evolution of the central star from the earliest
protostellar stage up to the early phases of nuclear burning on
the zero age main sequence by solving the conventional stellar
structure equations, taking into account mass accretion. The material
functions for the stellar evolution calculation -- the opacity,
the equation of state and other thermodynamic relationships,
as well as the nuclear reaction network -- assume zero metallicity.
The calculated evolution provides the emitted radiation
from the sink cell as a function of time.
We describe below our methods of calculation for the accretion flow
and the protostar, respectively.
\subsection{Radiation Hydrodynamics of the Accretion Flow}
To study the evolution of the flow of primordial gas we employ
a grid-based axisymmetric radiation hydrodynamic
code with self-gravity, previously used for studying
present-day star formation {\it (7,10)} and the evolution of
photoionized gas flow from protostellar disks
\cite{YW96,RY00}.
This code makes use of a nested-grid technique to cover a wide
spatial range \cite{YK95}. We have added a chemical network
besides changes based on the nature of the primordial material.
The governing equations for gas dynamics in cylindrical
coordinate $(R,Z)$ are,
\begin{equation}
\frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \mbox{\boldmath$v$}) = 0 ,
\end{equation}
\begin{equation}
\frac{\partial (\rho \mbox{\boldmath$v$})}{\partial t}
+ \nabla \cdot ( \rho \mbox{\boldmath$v$} \otimes \mbox{\boldmath$v$} ) =
- \rho \nabla \Phi - \nabla p + \frac{A^2}{\rho R^3} \mbox{\boldmath$n$}_R
+ \mbox{\boldmath$K$}
\end{equation}
\begin{equation}
\frac{\partial e}{\partial t} + \nabla \cdot (e \mbox{\boldmath$v$})
= - p \nabla \cdot \mbox{\boldmath$v$} + \Gamma - \Lambda ,
\end{equation}
\begin{equation}
p = (\gamma - 1) e ,
\end{equation}
where $\rho$ and $p$ are the gas density and pressure, $\Phi$
the gravitational potential,
$\mbox{\boldmath$v$}$ the 2-dimensional (2D) velocity vector $(v_R,v_Z)$,
$A \equiv \rho R v_\phi$ the angular momentum per unit volume,
$\mbox{\boldmath$n$}_R$ the radial unit vector,
$\mbox{\boldmath$K$}$ the radiation force,
$e$ the gas internal
energy density, $\Gamma$ and $\Lambda$ the heating and cooling
rates per unit volume, and $\gamma$ the adiabatic exponent.
We calculate $\gamma$ from the chemical composition
at each grid cell as in {\it (2)}.
The radiative and chemical processes included in
the energy source term $\Gamma - \Lambda$ are summarized in Table S1.
\begin{table}
\begin{center}
Table S1. Included thermal processes \\[3mm]
\begin{tabular}{llc}
\hline
\hline & Processes & References \\
\hline \\
& Photoionization & \cite{OF06} \\
Heating & Photodissociation & \cite{HM79} \\
& H$_2$ formation & \cite{HM79}, \cite{SK87} \\
\\
\hline
\\
& H$_2$ collisional excitation & \S~\ref{sssec:rad} \\
& H$^-$ free-bound emission & \S~\ref{sssec:rad} \\
& H$_2$ collisional dissociation & \cite{HM79}, \cite{SK87} \\
Cooling & H collisional ionization & \cite{OF06} \\
& H collisional excitation & \cite{An97} \\
& Compton scattering & \cite{An97} \\
& HeII collisional excitation & \cite{An97} \\
\\
\hline
\end{tabular}
\end{center}
\label{tab:thprocess}
\end{table}
\subsubsection{Radiation Transfer}
\label{sssec:rad}
For the problem at hand photons in different wavelength regimes
play vastly different roles in the thermal and chemical processes.
For example, whereas gas cools via molecular hydrogen line emission
and infrared/optical continuum radiation, photons with energies
$h \nu \geq 13.6$~eV (EUV) photoionize atomic hydrogen and
photons with energies $11.2~{\rm eV} \leq h \nu \leq 13.6~{\rm eV}$ (FUV)
photodissociate molecular hydrogen.
To treat their transfer properly
without time-consuming calculations, we adopt different
methods for each of the radiation components.
\paragraph{Molecular Hydrogen Line Cooling}
The primary cooling process in low-temperature
($< \sim 8000$K) primordial gas
is line emission via rotational and vibrational transitions
of hydrogen molecules.
We adopt the fitting formula by \cite{HM79} for the optically-thin
limit for $n < \sim 10^9~{\rm cm}^{-3}$.
Dense gas with $n > \sim 10^9~{\rm cm}^{-3}$ is opaque
to H$_2$ line emission, and the cooling rate is reduced
by photon trapping.
In this case the cooling rate is calculated by adapting
the method of \cite{Y06}
to the case of axial symmetry.
We sum up cooling rates for all possible transitions among the
rotational levels from $J=0$ to 20
and vibrational levels $v=0,1,2$.
The cooling rate by a transition for the optically thick case
is obtained by multiplying the value for optically thin
cooling by the escape probability.
The escape probability $\beta_{\rm esc}$ is evaluated as
\begin{equation}
\beta_{\rm esc} = \frac{\beta(\tau_R) + \beta (\tau_Z)}{2} ,
\end{equation}
where $\beta (\tau)$ is
\begin{equation}
\beta (\tau) = \frac{1 - \exp (\tau)}{\tau},
\end{equation}
and $\tau_R$ and $\tau_Z$ are the optical depths
of this transition along $R$ and $Z$ directions, respectively.
We calculate the optical depths using the local velocity gradients
(so-called Sobolev approximation)
\begin{equation}
\tau_{q} = \alpha \frac{c_s}{| \partial v_q / \partial q |} ,
\end{equation}
where $q=R$ or $Z$, $\alpha$ is the absorption coefficient
and $c_s$ is the sound speed.
For calculating the level-population of H$_2$ molecules, we omit the
excitation by absorbing stellar photons, which could potentially
reduce the line cooling rate.
In general, H$_2$ molecules could be excited to (i)
the Lyman-Werner bands by absorbing FUV photons
or (ii) the higher rotational and vibrational
levels by absorbing infrared photons.
In our case, the process (i) would be minor because the disk
is shielded against the stellar FUV radiation (also see Sec.~\ref{sec:disk}).
The process (ii) is negligible in most parts of the disk
even with the radiation from the protostar.
\paragraph{Continuum Cooling}
During the collapse phase before the formation of a protostar,
continuum cooling via H$_2$ collision-induced emission
(CIE) is important in the dense ($n > 10^{13}~{\rm cm}^{-3}$)
molecular gas {\it (2)}.
However, we only find such dense gas inside the sink cell;
it does not otherwise appear in our hydrodynamical calculation.
Instead, cooling via H$^-$ free-bound emission becomes
important in the nearly vertical flows onto the circumstellar
disk (also see Sec.~\ref{sec:disk} below).
The continuum cooling rate $\Lambda_c$ in the gray approximation
is written as
\begin{equation}
\Lambda_c = c \rho \kappa_P ( 4 \pi B(T_g) - E_c) ,
\end{equation}
where $c$ is the speed of light, and
$\kappa_P$ is the Planck mean opacity per unit mass.
The continuum radiation energy density $E_c$
satisfies the equation
\begin{equation}
\frac{\partial E_c}{\partial t}
= - \nabla \cdot \mbox{\boldmath$F$}_c + \Lambda_c + j_* ,
\label{eq:ecrad}
\end{equation}
where $\mbox{\boldmath$F$}_c$ is the energy flux, and $j_*$ is the stellar source term.
Using the total luminosity of the protostar $L_{\rm tot}$,
the stellar source term is given by
\begin{equation}
j_* = \frac{L_{\rm tot}}{2 \pi \Delta R^2 \Delta Z},
\end{equation}
where $\Delta R$ and $\Delta Z$ are the size of the sink cell,
which is equal to the cell size at the finest grid-level, 12~AU.
The source term is zero except for the sink cell.
We solve equations (\ref{eq:ecrad}) with the flux-limited
diffusion (FLD) approximation \cite{LP81} using
the operator splitting technique with $de/dt = - \Lambda_c$
\cite{TS01}.
The FLD method adopts a closure relation between $\mbox{\boldmath$F$}_c$
and $E_c$
\begin{equation}
\mbox{\boldmath$F$}_c = - \frac{c \lambda}{\rho \kappa_R} \nabla E_c ,
\label{eq:fldc}
\end{equation}
where $\kappa_R$ is the Rosseland mean opacity,
and $\lambda$ is the flux-limiter defined as
\begin{equation}
\lambda = \frac{2 + s}{6 + 3s + s^2} ,
\end{equation}
\begin{equation}
s = \frac{ | \nabla E_c |}{E_c \rho \kappa_R} .
\end{equation}
We use the opacities $\kappa_P$ and $\kappa_R$
for primordial gas calculated by \cite{MD05} in tabulated forms.
As mentioned above, the most important continuum cooling is
that via H$^-$ free-bound emission.
Neutral gas falling onto the disk is
heated up to $T_g \simeq 5 \times 10^3$~K by compressional heating,
until thermal balance is achieved with this cooling process.
In our calculation the accreting envelope is always
optically thin to the non-UV continuum radiation and non-UV stellar radiation
escapes without significant heating in the envelope.
\paragraph{Stellar EUV and FUV Radiation}
EUV radiation from the star ionizes the material in its immediate vicinity.
Within the HII region, there are two components of the EUV field:
the direct stellar component and the
diffuse component emitted via recombinations directly into the
ground state (i.e. no ``on-the-spot'' approximation).
We solve the transfer of these two EUV components separately
following \cite{RY00}.
We adopt a frequency-averaged approximation for each component,
taking into account the difference in their mean energies,
$h \bar{\nu}_*$ (stellar) and $h \bar{\nu}_d$ (diffuse).
To calculate the direct EUV field we
cast a number of radial rays from the central star to the outer edge of the
simulation box, along which the transfer equation for
the direct EUV photon number flux $\mbox{\boldmath${\cal F}$}_*$ is solved:
\begin{equation}
\nabla_r \cdot \mbox{\boldmath${\cal F}$}_* = - n (1 - x) \sigma_* \mbox{\boldmath${\cal F}$}_* ,
\label{eq:euvst}
\end{equation}
where $x$ is the degree of ionization and $\sigma_*$ is the absorption
cross section per particle.
The cross section $\sigma_*$ is a function of mean energy of
the direct EUV photons, i.e., the effective temperature of the star
$T_{\rm eff}$ \cite{OF06}.
We employ the FLD approximation for the diffuse component, whose
transfer equation is
\begin{equation}
\frac{\partial {\cal N}_d}{\partial t} =
- \nabla \cdot \mbox{\boldmath${\cal F}$}_d + \alpha_1 (T_g) n^2 x^2
- n (1-x) \sigma_d c {\cal N}_d ,
\label{eq:euvd}
\end{equation}
where ${\cal N}_d$ is the diffuse EUV photon number density,
$\alpha_1$ is the recombination coefficient to the ground state,
and $\sigma_d$ is the absorption cross section.
Under the approximation that the mean energy of the diffuse component
depends only on the local gas temperature $T_g$, the cross
section $\sigma_d$ becomes a function of $T_g$ \cite{OF06}.
In the FLD approximation,
\begin{equation}
\mbox{\boldmath${\cal F}$}_d = - \frac{c \lambda}{n (1-x) \sigma_d} \nabla {\cal N}_d ,
\end{equation}
in analogy to equation (\ref{eq:fldc}).
The photoionization heating and recombination cooling rates are
calculated for each component of EUV radiation separately.
As for the photodissociating FUV radiation,
we only consider the stellar direct component.
We calculate the photodissociation rate by multiplying
the value for the optically-thin case by the self-shielding factor
\cite{DB96}
\begin{equation}
f_{\rm sh} (N_{\rm H_2})= \left\{
\begin{array}{ll}
1 & (N_{\rm H_2} < N_1) \\
(N_{\rm H_2}/N_1)^{-3/4} &
(N_1 < N_{\rm H_2} < N_2) \\
0 & (N_{\rm H_2} > N_2) ,
\end{array}
\right.
\label{eq:fsld}
\end{equation}
where $N_{\rm H_2}$ is the H$_2$ column density
from the protostar to the point under consideration,
$N_1 = 10^{14}~{\rm cm}^{-2}$, and $N_2 = 10^{22}~{\rm cm}^{-2}$.
Although the original functional form of
$f_{\rm sh}(N_{\rm H_2})$ by \cite{DB96} is for the range
$N_1 < N_{\rm H_2} < N_2$,
we adopt the cut-off at $N_{\rm H_2} > N_2$,
where the photodissociation rate falls sharply
according to \cite{DB96}.
In our calculation, the column density largely exceeds
$N_2$ along the radial rays incident on the circumstellar disk.
As shown in Sec.~\ref{sec:disk} below, the innermost part of the disk
at $R <$ several 10~AU is not spatially resolved in our
calculations.
We expect the innermost part of the disk to be optically thick
and shield the material near the disk mid-plane behind it
from the stellar radiation \cite{TM04}.
We adopt the following treatment for modeling this effect.
During the calculation, we evaluate the disk scale height
$H (R) \equiv c_s/\Omega$ at each radius $R$ on the equator.
We compute $i_{R,d}$, the $R$-index of the cell at the outer edge
of the unresolved part, for which $H (R) < \Delta Z$.
The unresolved part is assumed to be opaque to radial rays incident
on the $(i_{R,d}+1)$-th cell on the equator up to a height $\Delta Z$,
but transparent for the other rays.
\paragraph{Radiation Force}
The total radiation force $\mbox{\boldmath$K$}$ is calculated as the sum
of contributions by the continuum radiation,
direct and diffuse EUV radiation:
\begin{equation}
\mbox{\boldmath$K$} = \frac{\rho \kappa_R}{c} \mbox{\boldmath$F$}_c
+ \frac{n (1-x) \sigma_*}{c} h \bar{\nu}_* \mbox{\boldmath${\cal F}$}_*
+ \frac{n (1-x) \sigma_d}{c} h \bar{\nu}_d \mbox{\boldmath${\cal F}$}_d .
\end{equation}
Here, we omit contributions from FUV photons.
In reality, the gas accretion envelope is opaque against
Lyman-$\alpha$ photons emitted from the stellar atmosphere and
HII region. Radiation pressure via the Lyman-$\alpha$ scattering
is exerted on the accretion envelope.
With mass accretion through the circumstellar disk, in particular, photons
are preferentially transferred toward the polar direction, where the gas
density rapidly decreases as the mass accretion proceeds
(``flashlight effect'': c.f. {\it (7)}). The semi-analytic modeling
by {\it (8)} shows that the Lyman-$\alpha$ pressure could influence the
dynamics of infalling material near the rotation axis.
As also discussed in {\it (8)}, however, the radiation pressure
via Lyman-$\alpha$ scattering would be significantly reduced once
gas in polar directions is blown away and photons
escape from the cavity. In this paper, we focus on the UV radiative
feedback effects to derive upper limits of the stellar final masses.
\subsubsection{Chemical Reactions in the Primordial Gas}
We consider the five species of primordial gas
H, H$^+$, e, H$_2$, and H$^-$,
based on the minimal model of \cite{Abel97} with some
additional reactions.
The included reactions are summarized in Table S2.
Except H$^-$, we solve kinetic equations
with an implicit difference scheme.
H$^-$ is assumed to be in chemical equilibrium.
Photodissociation of H$^-$ is omitted in the adopted
chemical network.
In our calculations the gas density is high enough that the
collisional destruction processes (R4, 14, and 15) control
its abundance. We curtail helium chemistry by
assuming that helium is singly ionized in an HII region
and atomic elsewhere.
We do not include deuterium reactions because HD
cooling is relevant only in low-temperature ($T_g < 200$~K)
and low-density ($n < 10^8{\rm cm^{-3}}$) gas,
which does not appear in our simulations.
\begin{table}[t]
\begin{center}
Table S2. Included chemical reactions \\[3mm]
\begin{tabular}{llc}
\hline
\hline No. & Reactions & References \\
\hline \\
R1 & H + e $\rightarrow$ H$^+$ + 2 e & \cite{Abel97} \\
R2 & H$^+$ + e $\rightarrow$ H + $\gamma$ & \cite{OF06} \\
R3 & H$^-$ + H $\rightarrow$ H$_2$ + e & \cite{GP98} \\
R4 & H$_2$ + H$^+$ $\rightarrow$ H$_2^+$ + H & \cite{GP98} \\
R5 & H$_2$ + e $\rightarrow$ 2 H + e & \cite{GP98} \\
R6 & H$_2$ + H $\rightarrow$ 3 H & \cite{SK87} \\
R7 & 3 H $\rightarrow$ H$_2$ + H & \cite{PSS83} \\
R8 & 2 H + H$_2$ $\rightarrow$ 2 H$_2$ & \cite{PSS83} \\
R9 & 2 H$_2$ $\rightarrow$ 2 H + H$_2$ & \cite{PSS83} \\
R10 & H + e $\rightarrow$ H$^-$ + $\gamma$ & \cite{GP98} \\
R11 & 2 H $\rightarrow$ H$^+$ + e + H & \cite{PSS83} \\
R12 & H + $\gamma$ $\rightarrow$ H$^+$ + e & \cite{OF06} \\
R13 & H$_2$ + $\gamma$ $\rightarrow$ 2 H
& \cite{TH85} and (\ref{eq:fsld}) \\
R14 & H$^-$ + e $\rightarrow$ H + 2 e & \cite{Abel97} \\
R15 & H$^-$ + H$^+$ $\rightarrow$ 2 H & \cite{GP98} \\
\\
\hline
\end{tabular}
\end{center}
\label{tab:reactions}
\end{table}
\subsubsection{Angular Momentum Transport in the Accretion Disk}
\label{sssec:ang}
In an exactly axisymmetric system, angular momentum must be conserved.
In reality, however, a massive circumstellar disk
can develop a non-axisymmetric spiral pattern,
which exerts a torque on the matter in the disk
and transfers angular momentum outward.
Recent 3-dimensional (3D) numerical simulations demonstrated that
this mechanism operates in fact in the circumstellar disks of the
first stars \cite{SGB10,Cl11}.
We mimic this effect by adopting the angular
momentum transport via the so-called $\alpha$-viscosity \cite{SS73}.
The equation of angular momentum transport is thus given by
\begin{equation}
\frac{\partial A}{\partial t} + \nabla \cdot (A \mbox{\boldmath$v$})
= - \frac{1}{R} \frac{\partial}{\partial R}
\left( R^3 \eta \frac{\partial \Omega}{\partial R} \right),
\end{equation}
where $\Omega$ is angular velocity,
$\eta = 2 \alpha \rho c_s^2 / ( 3 \Omega )$,
and $\alpha$ is a dimensionless free parameter.
We assume that the $\alpha$-parameter depends on the height from the equator,
\begin{equation}
\alpha(R,Z) = \alpha_0 \exp \left( - \frac{Z}{H (R)} \right) ,
\label{eq:alpha}
\end{equation}
where $\alpha_0$ is a constant.
The effective values of the $\alpha$-parameter in rapidly accreting
circumstellar disks are $\alpha \simeq 0.1 - 1$ as estimated from
3D simulations of present-day massive star formation
\cite{Krum09} as well as the formation of the first stars {\it (28)}.
In the fiducial case explained in the main article, we adopted
$\alpha_0 = 0.6$.
Figure \ref{fig:xmdot_a} shows the evolution of accretion rates
onto the protostar for different values of $\alpha_0$.
Although the evolution is qualitatively similar in all cases,
with higher $\alpha_0$ angular momentum is transferred more rapidly
and therefore accretion rates onto the protostar become higher.
We see that the final stellar mass increases with $\alpha_0$.
With the low value of $\alpha_0 = 0.3$, the final stellar mass is
about $35~M_\odot$. This is close to the low-mass limit of $30~M_\odot$
predicted by {\it (3)}, who expected that this amount of gas would
accrete onto the protostar in a few thousand years, which is much
shorter than the timescale of the protostellar evolution.
Even with $\alpha_0$ as large as unity, on the other hand,
the final mass reaches at most $M_{\ast} \simeq 50~M_\odot$.
Note, however, that this does not exclude the possibility of
formation of higher-mass stars.
The semi-analytic models predict that more massive stars would
form with weaker initial rotation of the natal core {\it (8)}
(also see Sec.~\ref{sec:mt08}).
Although rare, stars exceeding $100~M_\odot$ might still form
in such circumstances.
Figure \ref{fig:xmdot_a} also shows the evolution in such
a test case, whereby the initial angular momentum is artificially
reduced to 30\% of the original
value.\footnote{The angular momentum is reduced at the beginning of
the 2D calculation, e.g., at a point when the central density is
$\simeq 10^6~{\rm cm}^{-3}$ in the run-away collapse
stage (see Sec.~\ref{sec:setup}).}
The final stellar mass is $\simeq 85~M_\odot$ in this case.
The increase of the final mass is understood as follows.
First, the gas density remains high in the accretion envelope in the
polar directions for the same stellar mass.
The stellar EUV photons are consumed more effectively by photoionization
of neutral hydrogen generated via rapid recombination, which delays growth
of an HII region.
Second, the protostellar evolution differs from the fiducial
case, because of the higher accretion rates.
Figure \ref{fig:pevol_a} shows that, with the weaker rotation,
the Kelvin-Helmholtz contraction stage is shifted to higher
stellar masses. Correspondingly, the stellar EUV luminosity increases
rapidly at a stage when the protostar is more massive than for the
fiducial case.
Because of these effects, the stellar UV feedback becomes effective
at higher stellar masses.
We note that for all examined cases, the mass accretion ceases
soon after the protostar's arrival to the zero age main sequence.
\subsection{Protostellar Evolution}
We follow the evolution of the central protostar by solving
the four stellar structure equations taking account of
mass accretion \cite{OP03, HO09, HYO10}:
\begin{equation}
\left( \frac{\partial r}{\partial M} \right)_t = \frac{1}{4 \pi \rho r^2},
\label{eq:con}
\end{equation}
\begin{equation}
\left( \frac{\partial P}{\partial M} \right)_t = - \frac{GM}{4 \pi r^4},
\label{eq:mom}
\end{equation}
\begin{equation}
\left( \frac{\partial L}{\partial M} \right)_t
= \epsilon - T \left( \frac{\partial s}{\partial t} \right)_M ,
\label{eq:ene}
\end{equation}
\begin{equation}
\left( \frac{\partial s}{\partial M} \right)_t
= \frac{G M}{4 \pi r^4} \left( \frac{\partial s}{\partial p} \right)_T
\left( \frac{L}{L_s} - 1 \right) C ,
\label{eq:heat}
\end{equation}
where $M$ is the Lagrangian mass coordinate, $\epsilon$ is the energy
production rate by nuclear fusion, $s$ is the specific entropy,
and $L_s$ is the radiative luminosity with adiabatic temperature
gradient.
The coefficient, $C$ in equation (\ref{eq:heat}) is unity if
$L < L_s$ (i.e., in radiative layers), and given by the
mixing-length theory if $L > L_s$ (i.e., in convective layers).
We also solve the structure of the accretion flow
inside the sink cell under the assumption of
steady state and spherical symmetry.
The entire structure of both the protostar and accretion flow
is consistently determined to satisfy the jump
conditions for the accretion shock at the stellar surface \cite{SST80}.
Mass accretion rates onto the protostar $\dot{M}_*$ are given
by mass inflow rates through the surface of the sink cell
in the radiation-hydrodynamics calculation.
With a high accretion rate of
$\dot{M}_* > 10^{-4}~M_\odot~{\rm yr}^{-1}$,
the accretion flow just above the stellar surface becomes opaque
to the stellar radiation under the assumption of perfect spherical symmetry.
The photosphere is located far from the stellar surface,
its temperature is reduced to $\simeq 6000$~K
even for very massive stars of $M_* \simeq 100~M_\odot$ {\it (11)}.
With the realistic disk accretion, however,
the stellar surface at high latitude is not totally embedded in such
an opaque flow and the UV radiation can freely radiate.
For this reason, we evaluate the protostellar EUV and FUV photon number
luminosities using the fluxes at the stellar surface as
\begin{equation}
S_{\rm EUV} = 4 \pi R_*^2 \int_{\rm 13.6eV}^\infty
\frac{\pi B(T_{\rm eff})}{h \nu}~d \nu ,
\label{eq:sev}
\end{equation}
\begin{equation}
S_{\rm FUV} = 4 \pi R_*^2 \int_{\rm 11.2eV}^{\rm 13.6eV}
\frac{\pi B(T_{\rm eff})}{h \nu}~d \nu ,
\end{equation}
where $R_*$ is the stellar radius, $B(T_{\rm eff})$ is the Planck
function, and $T_{\rm eff}$ is defined as
\begin{equation}
T_{\rm eff} = \left( \frac{L_* + L_{\rm acc}}{4 \pi \sigma R_*^2}
\right)^{1/4} ,
\label{eq:teff}
\end{equation}
where $L_*$ is the stellar luminosity,
$L_{\rm acc} \equiv G M_* \dot{M}_* / R_*$ is the accretion luminosity,
and $\sigma$ is Stefan-Boltzmann constant.
The stellar UV luminosities given by equations (\ref{eq:sev})
- (\ref{eq:teff}) assume that all the gas reaching the stellar
surface releases its gravitational energy at the stellar surface.
In reality, some fraction of the gravitational energy
would be radiated away from the circumstellar disk with a
lower effective radiation temperature than that of
the stellar surface.
Because the accretion luminosity $L_{\rm acc}$ is
only a few $10$~\% of the stellar luminosity $L_*$, when
the protostellar feedback begins to influence the mass accretion
(see the main article), we do not expect a significant change in the overall
accretion process.
The stellar EUV photons are consumed mostly by photoionizing the recombined
hydrogen atoms within the HII region. As the recombination rate is
proportional to the square of density, this consumption is most
significant in the vicinity of the protostar.
In our calculations, however, the flow structure very near the
protostar is masked by our assumption of a sink cell.
We evaluate the EUV consumption rate within the sink cell
as follows.
First, we consider the spherical ``evacuation zone'', whose radius
$R_{\rm ev}$ is smaller than the gravitational radius for the ionized gas,
\begin{equation}
R_{g,{\rm HII}} \equiv \frac{G M_*}{c_{s,{\rm HII}}^2}
\simeq 100~{\rm AU} \left( \frac{M_*}{10~M_\odot} \right)
\left( \frac{T_{\rm HII}}{10^4~{\rm K}} \right)^{-1} ,
\label{eq:rghii}
\end{equation}
and larger than the size of the sink cell, $\simeq 10$~AU.
In our calculations the temperature of ionized gas $T_{\rm HII}$ is
$\simeq 3 \times 10^4~K$ just after formation of the HII region and
rises slightly as the stellar mass increases.
We adopt $R_{\rm ev} = 30$~AU as a fiducial value.
The density distribution within the evacuation zone should be well
approximated as the free-fall flow whose radial density structure
follows $\rho(r) \propto r^{-3/2}$. The consumption rate of EUV
photons within this zone is analytically written as,
\begin{equation}
S_{\rm EUV, ev} = \frac{\alpha \dot{M}_{\rm ev}^2}{8 \pi \mu^2 G M_*}
\ln \left( \frac{R_{\rm ev}}{R_*} \right) ,
\label{eq:seuv}
\end{equation}
where $\alpha$ is the total recombination rate,
$\mu = (1 + 4 y_{\rm He}) m_p$ with the helium abundance $y_{\rm He}$
and proton mass $m_p$, and $\dot{M}_{\rm ev}$
is the mass inflow rate into this zone.
We evaluate $\dot{M}_{\rm ev}$ using the gas inflow rate
along the $Z$-axis,
\begin{equation}
\dot{M}_{\rm ev} = 4 \pi \rho_{z,{\rm ev}} R_{\rm ev}^2
\sqrt{ \frac{2 G M_*}{R_{\rm ev}} } ,
\end{equation}
where $\rho_{z,{\rm ev}}$ is the gas density at $(R,Z) = (0,R_{\rm ev})$.
We suppose that the HII region is quenched inside the evacuation
zone when the stellar EUV luminosity $S_{\rm EUV}$ is less than the
consumption rate $S_{\rm EUV,ev}$.
We do not solve the EUV radiative transfer for such cases.
After $S_{\rm EUV}$ exceeds $S_{\rm EUV,ev}$, we remove the limit
of $S_{\rm EUV,ev}$ and assume that the EUV luminosity from the sink
is $S_{\rm EUV}$.
We only marginally resolve the evacuation zone with the current
grid resolution.
The estimated value of $S_{\rm EUV,ev}$ depends on
the grid size without resolving $R_{g,{\rm HII}}$,
because free-fall flow is valid only inside of $R_{g,{\rm HII}}$.
For test calculations with a 2$\times$ coarser innermost grid around the
protostar, formation of the HII region begins for a bit higher
stellar mass (by a few $10\%$).
Moreover, the flow structure within the evacuation zone could be
complex. For instance, the protostellar outflow might be launched from
the innermost part of the disk \cite{Mcd06} with the help of magnetic field
generated by dynamo amplification \cite{TB04, SL06}.
This outflow would help the breakout
of the HII region by clearing out materials close to the star
in the polar directions. However, the evolution should depend on the
detailed density structure in the outflow-launching region,
which controls the EUV consumption rate.
\section{Simulation Setup}
\label{sec:setup}
As the initial condition of our calculation, we assume the structure of
a dense core in the run-away collapse from the cosmological simulation
by {\it (6)}.
The calculation by {\it (6)} followed the entire evolution
from the cosmological initial condition to the birth of a
primordial protostar under the standard $\Lambda$CDM cosmology.
A small protostar of $M_* \simeq 0.01~M_\odot$ forms
at $10^{20}~{\rm cm}^{-3}$ as a result of the run-away collapse of a
dense primordial-gas core at the cosmological redshift $z = 14$.
Specifically, we take the central 0.3~pc cube around the density peak
when the maximum density is $10^6~{\rm cm}^{-3}$ as our initial condition.
We reduce the 3D data to an axisymmetric 2D distribution by averaging
over azimuthal angles. Our simulation box contains the total gas mass
of $\simeq 300~M_\odot$.
The numbers of the grid cells are initially
$N_Z \times N_R = 42 \times 42$, including 2 ghost cells in each direction.
After we start our axisymmetric calculation, the dense core
experiences continued gravitational collapse.
We increase the grid resolution for the central collapsing
region by successively adding finer nested grids as needed.
With this nested-grid technique, we always resolve the minimum
Jeans length by tens of grid cells.
The increase of the grid-level is limited up to 8 owing to
computational cost for following the subsequent accretion phase
until the final stellar mass is fixed.
We terminate the calculation of the run-away collapse when
the minimum Jeans length becomes too short to be resolved with
the finest grids.
At this point we create a sink cell at the origin
and calculate the subsequent accretion phase as described in Sec.~1.
The central density at this moment is $\simeq 10^{12}~{\rm cm}^{-3}$,
and the cell size at the finest grid-level is $\simeq 12$~AU.
This recipe enables us to smoothly connect the 3D cosmological
simulation using the particle method to a 2D (axial symmetry assumed)
local simulation using the nested-grid method.
We have confirmed that the evolution in our 2D calculation
is reasonably consistent with the 3D results after the maximum
density exceeds $10^6~{\rm cm}^{-3}$.
For example, the upper panel of Figure~\ref{fig:fkep_kprm} shows a
comparison of the angular-momentum profiles for
the 3D and 2D calculations using
a parameter,
\begin{equation}
\label{eq:fkep}
f_{\rm Kep} (M_r) \equiv \frac{V_{\phi,r}}{V_{\rm Kep,r}}, \qquad
V_{\phi,r} = \frac{l_r}{r}, \
V_{{\rm Kep},r} = \sqrt{ \frac{G M_r}{r} },
\end{equation}
where $M_r$ is the enclosed mass within radius $r$, and
$l_r$ is specific angular momentum averaged over the spherical shell
whose radius is $r$.
The profile for the 2D simulation is within 20~\% offset from that
in the 3D case taken from {\it (6)}.
Although most of our simulations were done with the settings described
above, we also calculated early evolution, turning off the stellar feedback,
with the higher maximum grid-level of 9.
The spatial resolution for $R < 240$~AU is doubled in this case and
size of the finest cell is $\simeq 6$~AU.
Figure \ref{fig:xmdot_a} shows the evolution of the accretion rate
until the stellar mass reaches $30~M_\odot$.
The accretion rates are initially a bit lower with the higher central
resolution but converge to our fiducial case as the stellar
mass increases.
\section{Structure of the Circumstellar Disk}
\label{sec:disk}
Here, we examine the detailed structure of the circumstellar
disk observed in our fiducial case.
Figure \ref{fig:1000au_snapshot} shows the 2D structure of the
region within 1000~AU of the protostar, when the stellar mass is
$\simeq 10~M_\odot$ and $20~M_\odot$.
The snapshot for the $10~M_\odot$ star shows the structure before
the formation of an HII region.
We see that the accreting gas hits the disk surface and
heats up to $\simeq 4000$~K by compression.
Radiative cooling via H$^-$ free-bound emission operates here
as explained in Sec.~\ref{sssec:rad}.
Figure \ref{fig:disk} displays the radial structure of the disk
at this moment.
The hydrogen is primarily in molecular form within the
disk due to rapid formation via the gas phase three-body reaction.
Note that the existence of H$_2$ molecules and their contribution for
radiative cooling are not included in the semi-analytic models.
The total mass of this molecular gas is $\simeq 10~M_\odot$, which
is comparable to the stellar mass.
Gas on the equator at $R < 1000$~AU has the rotational velocity more
than 80\% of the local Kepler value.
The disk scale height is resolved
on average except for the innermost part at $R < 100$~AU.
The profile of the Toomre Q-parameter, a measure of the
gravitational stability of the disk, shows that the minimum
value $Q_{\rm min}$ is almost unity.
Thus, the disk is marginally gravitationally stable against
the fragmentation.
However, high-resolution 3D simulations demonstrate that
the disk fragments due to gravitational instability in
very early stage of the accretion phase {\it (28)}.
Note that ``ring-like'' fragmentation can be captured even in
axisymmetric 2D simulations.
The absence of fragmentation in our simulation is probably due to
the limited spatial resolution and the large values
of the viscous $\alpha$-parameter, $\alpha_0 \geq 0.3$
(also see Sec.~\ref{sssec:ang}).
In our case, the disk is also marginally stable in an earlier
stage when the stellar mass is $\simeq 3~M_\odot$ (Fig.~\ref{fig:disk}).
This is almost the same even with the doubled spatial resolution for
$R < 240$~AU.
Even disks which experience fragmentation have a quasi-static structure
with $Q_{\rm min} \simeq 1$, if averaged over many rotation periods.
Such structure is well mimicked in our calculations.
The right panel in Figure~\ref{fig:1000au_snapshot} shows the
2D disk structure just after the birth of the HII region.
We see a strong photoevaporating flow in the polar directions within
the HII region. The stellar FUV luminosity has also increased
significantly, but the H$_2$ molecular disk still exists, shadowing the
stellar photodissociating photons from the dense equatorial regions,
evident in the disk's temperature and density distributions. The mass
of the molecular gas is $\simeq 10~M_\odot$ at this moment.
\section{Comparison to Semi-analytic Models}
\label{sec:mt08}
Our calculations show that mass accretion toward forming
first stars is shut off via dynamical expansion of the HII region and
photoevaporation of the circumstellar disk.
This is qualitatively consistent with the picture predicted by
the semi-analytic models {\it (8)}.
Their models show that the final stellar mass is
$\simeq 145~M_\odot$ for their fiducial case; this value varies
with different sets of input parameters.
One of the input parameters is $f_{\rm Kep}$ given by equation
(\ref{eq:fkep}); their adopted fiducial value is $0.5$.
Figure~\ref{fig:fkep_kprm} shows that
$f_{\rm Kep} \simeq 0.6-0.7$ for our fiducial case and is
lower than $0.5$ for the weak-rotation case.
For the semi-analytic model with $f_{\rm Kep} > 0.25$, however,
the final stellar mass does not depend strongly on $f_{\rm Kep}$,
but rather on the entropy of the accreting gas,
measured with a non-dimensional parameter
\begin{equation}
K' \equiv \frac{P / \rho^\gamma}{1.88 \times 10^{12} {\rm cgs}}
= \left( \frac{T}{300~{\rm K}} \right)
\left( \frac{10^4 {\rm cm}^{-3}}{n} \right)^{0.1} ,
\end{equation}
where $\gamma = 1.1$ is adopted for a typical value for
the primordial gas {\it (2)}.
The semi-analytic models use $K'=1$ for their fiducial case and
show that the final mass roughly scales as $K'^{1.3}$.
From Figure \ref{fig:fkep_kprm} we see that
$K' \simeq 0.7$ for our fiducial case and $K' \simeq 0.8 - 1$ for
the weak-rotation case.
The semi-analytic model with $K' = 0.7$ predicts a final mass
$\simeq 90~M_\odot$, twice our value of $45~M_\odot$.
The final mass $\simeq 85~M_\odot$ for our weak-rotation case is
somewhat lower than the semi-analytic prediction $> 108~M_\odot$
for $K' > 0.8$.
The initial conditions in our examined cases correspond
to typical values for gas clouds bearing the first stars.
Figure~\ref{fig:xmdot_a} shows that the accretion rates
for the cases considered lie in the range
$10^{-2}$ to $10^{-3}~M_\odot~{\rm yr}^{-1}$ for the
$10~M_\odot$ star, comparable to that
expected for 12 mini-halos using the semi-analytic models \cite{TSO10}.
However, our calculations attain systematically lower final
masses than predicted by semi-analytic models.
We examine possible reasons for this difference below.
First, in our calculations, the mass ratio between the star and disk
$f_d \equiv M_{\rm disk}/M_*$ differs from the fiducial value in the
semi-analytic models, $f_d = 0.3$.
As discussed in Sec.~\ref{sec:disk}, for our fiducial case the
mass of the molecular disk is $\simeq 10~M_\odot$, when the
stellar mass is $10~M_\odot$, and would be even higher, if we
include the outer atomic part.
In 2D simulations, however, the value of $f_d$ depends on
the degree of $\alpha$-viscosity, $\alpha_0$ (see Sec.~\ref{sssec:ang}).
With the larger value $\alpha_0 = 1$, for example, the mass of the
molecular disk for the $10~M_\odot$ star is reduced to $8~M_\odot$.
To derive the appropriate value of $f_d$, we need detailed
3D numerical simulations which solve the transport of angular
momentum in self-gravitating disks.
Recent work on the Galactic star formation demonstrates that
massive disks with $f_d \geq 1$ can form in some situations, though
the value of $f_d$ varies with different initial conditions
of gas cores \cite{Mcd10,Kratter10}.
We contend, however, that differences in the stellar feedback process
are principally responsible for the differences in final stellar mass.
The semi-analytic models adopt the following formula for estimating
the photoevaporation rate of the disk,
\begin{equation}
\dot{M}_{\rm evp} = 4.1 \times 10^{-5}
~\left( \frac{S_{\rm EUV}}{10^{49}~{\rm sec}^{-1}} \right)^{1/2}
\left( \frac{T_{\rm HII}}{10^4~{\rm K}} \right)^{2/5}
\left( \frac{M_{\rm *d}}{100~M_\odot} \right)^{1/2}
\ \ M_\odot~{\rm yr}^{-1} ,
\label{eq:xmdot_evp}
\end{equation}
where $M_{* d} = M_* + M_{\rm disk}$ \cite{HJLS94}.
For our fiducial case the photoevaporation rate is about
$2 \sim 3\times 10^{-4}~M_\odot~{\rm yr}^{-1}$, after the stellar mass
exceeds $30~M_\odot$.
By contrast, equation (\ref{eq:xmdot_evp}) estimates an evaporation
rate of several $10^{-5}~M_\odot~{\rm yr}^{-1}$,
using $T_{\rm HII} = 5 \times 10^4$~K and the values of $S_{\rm EUV}$
and $M_{\rm *d}$ from the simulation ($M_{\rm disk}$
is the mass of the molecular disk).
We attribute this difference to the basic assumption of an infinitely
thin disk for
deriving equation (\ref{eq:xmdot_evp}) \cite{HJLS94}.
As a result, only the diffuse EUV radiation (see Sec.~\ref{sssec:rad})
is accounted for in equation (\ref{eq:xmdot_evp}).
In our calculations the stellar direct EUV radiation field impinging on
the flared disk is primarily responsible for determining
the photoevaporation rate.
Even if the transfer of diffuse EUV radiation is turned off
in the simulation, we do not see any significant change in
the evolution.
Moreover, the semi-analytic models assume that,
after the HII region forms,
mass accretion onto the circumstellar disk still continues
from regions shaded by the disk, where the stellar UV radiation is blocked.
In this picture the stellar final mass is determined when
the photoevaporation rate exceeds the mass
accretion rate from the shaded regions.
This does not agree with our simulations.
It is true that the stellar UV radiation is blocked behind
the disk as shown in Figure 2 in the main article.
In our calculations, however, a shock front propagates into the
accretion envelope shaded by the disk, as the HII region
dynamically expands.
Figure~\ref{fig:shots_pn} shows that there is the outward pressure gradient
within the HII region due to the photoevaporating flow.
As the HII region dynamically expands, the shocked
accretion envelope behind the disk also obtains the same
pressure gradient.
As a result, the shocked gas is accelerated outward due to this
pressure gradient at a velocity of several km/sec.
This is shown in Figure \ref{fig:vrz0}. We see that the gas
at $R >$ a few $10^3$~AU is moving outward when the
stellar mass exceeds $20~M_\odot$.
The outflow rate of atomic gas via this horizontal motion is
$9 \times 10^{-4}~M_\odot~{\rm yr}^{-1}$, when the stellar mass
is $35~M_\odot$.
The mass supply from the envelope to the disk is cut off by this
moment.\footnote{In our simulations, we adopt a
semi-permeable outer boundary condition, which allows outflow
but prohibits mass inflow through the boundary.
In principle this boundary condition can ultimately limit mass
accretion onto the disk. However, mass infall toward the disk is
reversed and the gas flows outward through the outer boundary
long before this limit is reached.}
The isolated disk without replenishment from the envelope continues
to lose gas by the photoevaporation. Ultimately, mass accretion
onto the protostar ceases and the stellar final mass
is fixed.
\section{Potential Effects of 3-Dimensionality}
For this study we have adopted an axisymmetric 2D code for
calculating the dynamical evolution of the accreting flow.
Recent 3D simulations for the present-day high-mass star
formation are helpful for speculating on potential 3D effects.
A primary radiative feedback effect from Galactic high-mass stars
is radiation pressure exerted on dust grains coupled with the gas
accretion flow.
Historically, this effect was first studied under spherical symmetry
\cite{Kahn74,YK77,WC87} and it was
found that the accretion flow toward high-mass stars is prevented
by the strong repulsive force (called as "radiation pressure problem").
2D numerical simulations show that non-spherical mass accretion
via disks significantly alleviates this problem \cite{YS02,kuiper10}.
3D simulations by \cite{Krum09} demonstrate that 3D effects
(e.g., radiative Rayleigh-Taylor instability growing in the
accretion envelope) also reduce the repulsive force in addition
to the effect of the disk accretion.
However, it is still controversial whether the 3D effects are
essential for determining the upper mass limit for
present-day high-mass star formation.
More recent 3D simulations by \cite{kuiper11} show that the radiation
pressure barrier is circumvented with the disk accretion, but find no
critical 3D effects.
Without dust grains the opacity of the primordial gas is much lower
than for the present-day Galactic interstellar medium.
The dynamical evolution discussed in our article is
not primarily due to radiation pressure, but rather to the
high gas pressure of photoionized gas, which causes the dynamical
expansion of the HII region supplemented by the photoevaporation
of the disk.
UV feedback effects in the present-day high-mass star forming regions
have also been studied with 3D numerical simulations \cite{Peters10a, Dale05}.
However, these studies focus on the feedback effects over the large
length-scale of cluster-forming clumps, $\sim 1$~pc.
By contrast, we consider here the smaller-scale feedback processes in the
vicinity of protostars, such as photoevaporation of the disks.
The photoevaporation rate from the disk is expected to be highest at a radial
distance from the star $R = R_{g,{\rm HII}}$ given by equation
(\ref{eq:rghii}) \cite{HJLS94}.
However, the regions smaller than $1000$~AU around forming stars are
not spatially resolved in the above-cited studies.
The photoevaporation of the disk is a local process which
takes place in the very central part of the HII region.
In 3D we can expect the photoionized gas
to have a clumpier structure than modeled by our 2D simulations.
However, such small-scale inhomogeneities would be smoothed out
within the sound crossing time, which is much shorter than the dynamical
timescale of the expanding HII region.
As referred in Sec.~\ref{sssec:ang}, recent 3D simulations
demonstrate that the circumstellar disks around primordial
protostars fragment in the early stage of the accretion phase.
The fragmentation of the disk would somewhat reduce the stellar final
masses, as the accreting materials are shared among the fragments
\cite{Peters10b}.
Several authors have also suggested that a turbulent velocity field is
present after the gravitational collapse of primordial gas cores {\it (3)}.
Although the expected turbulent field is not strong for the very
first stars, we note that, in our case, the turbulent field is further
weaken by mapping the 3D data to the 2D axisymmetric data
(see Sec.~\ref{sec:setup}).
With a random velocity field the angular momentum directions
of accreting materials would be time-dependent. As a result, the
polar direction of the circumstellar disk should vary with time as well
and the photoevaporating flow would clear out a larger amount
of accreting materials than otherwise.
We speculate that this effect would accelerate the horizontal
expansion of the HII region and reduce the resulting final stellar mass.
Our obtained final mass of the first stars,
several $10~M_\odot$, is thus a conservative upper limit.
Nonetheless, this is still much lower than the postulated high
values exceeding $100~M_\odot$.
\clearpage
\setcounter{figure}{0}
\renewcommand{\thefigure}{S\arabic{figure}}
\begin{figure}
\begin{center}
\psfig{file=xmdot_a.eps,width=0.9\textwidth}
\caption{Evolution of accretion rates onto the protostar
with different $\alpha$-parameters for angular momentum
transport. The blue, red, and green curves depict the
results with $\alpha_0 = 1$, 0.6, and 0.3 in equation
(\ref{eq:alpha}), respectively.
The magenta curves display the evolution with the initial
angular momentum reduced to 30\% of the fiducial value.
For each case the solid and dashed lines represent the evolution with
and without radiative feedback from the protostar,
respectively.
The red filled squares represent the no-feedback case with
$\alpha_0 = 0.6$ doubling the spatial resolution in the central
region of $R < 240$~AU (also see Sec.~\ref{sec:setup}).
}
\label{fig:xmdot_a}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=pevol_a.eps,width=0.9\textwidth}
\caption{Accretion rates onto the protostar as a function of
the accreted stellar mass, assuming
different $\alpha$-parameters for angular momentum
transport. The blue, red, and green curves depict the evolution
with $\alpha_0 = 1$, 0.6, and 0.3 in equation
(\ref{eq:alpha}), respectively.
The magenta curves display the evolution with the initial
angular momentum reduced to 30\% of the fiducial value.
For each case the solid and dashed lines represent the evolution with
and without radiative feedback from the protostar,
respectively.}
\label{fig:pevol_a}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=fkep_kprm.eps,width=0.9\textwidth}
\caption{Structure of the accreting envelope
at the end of the calculation of the run-away collapse stage for
the fiducial case (red lines) and weak-rotation case (magenta lines).
{\it Upper panel:} the ratio of the local rotational velocity
to Kepler velocity $f_{\rm kep}$.
The red open circles represent a snapshot at the birth of a embryo
protostar in a three-dimensional cosmological simulation {\it (6)}.
{\it Lower panel:} Dimensionless entropy of the
accreting gas $K'$. The mass-weighted average values $\overline{K'}$
are plotted against the enclosed gas mass.}
\label{fig:fkep_kprm}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=1000au_snapshot.eps,width=1.0\textwidth}
\caption{
The structure of gas temperature (left-hand side) and of number
density and velocity (right-hand side) in the vicinity of $1000$~AU
around the protostar.
The left and right panels show the snapshots when the stellar
mass is $\simeq 10~M_\odot$ and $20~M_\odot$.
The elapsed time since the birth of the protostar is also shown
in each panel.
Note that the color scale of the density, the legend of the velocity
vectors, and size of the plotted area are different from those in
Figure 2 in the main article.
}
\label{fig:1000au_snapshot}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=disk.eps,width=0.6\textwidth}
\caption{Radial structure of the circumstellar disk
within $1000$~AU around the protostar when
the stellar mass is $3~M_\odot$ (green lines for the lower two panels),
$10~M_\odot$ (red lines) and $20~M_\odot$ (blue lines).
{\it Upper panel:} The fraction of hydrogen molecules along
the equator $f_{\rm H_2}$ (solid lines), and the ratio
of the local rotational velocity to Kepler velocity
$f_{\rm Kep}$ (dashed lines).
{\it Middle panel:} The scale height of the disk $H$. The black
thin line shows the grid resolution in our standard cases, and the
discontinuity at $R \simeq 470$~AU indicates a grid-level boundary there.
The profile when the stellar mass is $3~M_\odot$
with the doubled spatial resolution for $R < 240$~AU
is also presented (green filled squares).
The black thin dashed line shows the grid resolution in this case.
{\it Lower panel:} Same as the middle panel but for
Toomre Q-parameter at each radius.
}
\label{fig:disk}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=snapshots_pn.eps,width=1.0\textwidth}
\caption{Same as Fig.~2 in the main article but for gas temperature
multiplied by number density in the left panel, which is
nearly proportional to gas pressure.
The left and right panels show snapshots at times, when the stellar
mass is $M_* = 32~M_\odot$ and $37.5~M_\odot$ respectively.
}
\label{fig:shots_pn}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\psfig{file=vrz0.eps,width=1.0\textwidth}
\caption{Distribution of $R$-component of the radial velocity $v_R$
along the equator ($Z=0$).
The snapshots when the stellar mass is $10~M_\odot$ (black),
$20~M_\odot$ (red), $32~M_\odot$ (blue), $37.5~M_\odot$ (magenta),
and $42~M_\odot$ (green) are plotted.
Positive $v_R$ means that gas is moving outward.
}
\label{fig:vrz0}
\end{center}
\end{figure}
\end{document}
|
\section{
\renewcommand\section{
\@startsection{section}{3}{\z@}%
{-3.25ex\@plus -1ex \@minus -.2ex}%
{1.5ex \@plus .2ex}%
{\normalfont\normalsize\bfseries\mathversion{bold}}}
\renewcommand\subsection{
\@startsection{subsection}{3}{\z@}%
{-3.25ex\@plus -1ex \@minus -.2ex}%
{1.5ex \@plus .2ex}%
{\normalfont\normalsize\bfseries\mathversion{bold}}}
\renewcommand\subsubsection{
\@startsection{subsubsection}{3}{\z@}%
{-3.25ex\@plus -1ex \@minus -.2ex}%
{1.5ex \@plus .2ex}%
{\normalfont\normalsize\itshape}}
\makeatother
\makeatletter \@addtoreset{equation}{section} \makeatother
\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
\makeatletter
\renewcommand{ |
\section{Introduction}
The Diffuse Ionized Medium (DIG) was detected through its optical line
emission outside the classical H~{\sc ii}\, regions
(\cite[Reynolds (1971)]{1971PhDT.........1R}) and turns out to be a
major component of the interstellar medium in galaxies (\cite[Reynolds
(1991)]{1991IAUS..144...67R}). Most
specialist agree that OB stars in galaxies likely represent the main
source of ionizing photons for the DIG (see \cite[Haffner et al. (2009)]{2009RvMP...81..969H}). However, an additional
ionizing source is suggested by reported increase of such emission
line ratios as [N~{\sc ii}]/H$\alpha$\xspace, [S~{\sc ii}]/H$\alpha$\xspace with galactic height. Several
sources of additional ionization/heating have been suggested without
complete success.
\section{Proposed Scenario}
The extraplanar DIG is destributed in clouds (for simplicity
represented by rectangles) that are ionized by two star populations:1)
OB stars located in the thin disk and whose ionizing radiation escape
from the disk through small ``holes'', which ionize the ``bottom''
part of the gas clouds. They constitute only a small fraction of the
entire population of the OB stars in the disk. 2) HOLMES, which are
distributed in the galaxy thick disk and halo. Their influence with
respect to that of OB stars increases away from the galacti
plane. Figure \ref{fig1} shows the datails of the proposed scenario.
\begin{figure}
\begin{center}
\includegraphics[width=2.1in]{fig-holmes2.eps}
\caption{Schematic representation of the extraplanar gas and ionizing
stars. Ionized gas is represented in blue, neutral gas in red, with
a lighter shade for less dense regions. For both components, a
darker shade indicates gas of higher densty. The big blue stars in
the plot represent OB stars. The small purple stars represent teh
HOLMES, which can be either central satrs of present-day planetary
nebulae, or hot pre-white dwarfs.}
\label{fig1}
\end{center}
\end{figure}
\section{Observational Data}
We chose to focus on the edge on spiral galaxy NGC 891, a
galaxy that has been extensively observed, especially in optical
emission lines (Figure \ref{fig2}), providing the best diagnostics for our
scenario.
\begin{figure}
\begin{center}
\includegraphics[width=3.in]{fig-holmes1.eps}
\caption{Observed values of [O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi, [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi and [N~{\sc ii}]/[O~{\sc ii}] in NGC
891, as a function of the distance to the galactic plane. The data
are from \cite[Otte et al. (2001)]{2001ApJ...560..207O}.}
\label{fig2}
\end{center}
\end{figure}
\section{Modeling}
The ionizing spectral energy distribution (SED) from OB stars is
obtained using the code Starburst99 (\cite[Leitherer et al. (1999)]{1999ApJS..123....3L}), considering a continuous star
formation. The SED of HOLMES is obtained using the code PEGASE (\cite[Fioc
\& Rocca-Volmerange (1997)]{1997A&A...326..950F})
considering a instantaneous starburst at look back after 10 Gyr. The
photoionization models for the DIG were computed with Cloudy (\cite[Ferland et al. (1998)]{1998PASP..110..761F}). Each
one is defined by the ratio of the surface fluxes: $\Phi_{\mathrm{HOLMES}}$\xspace / ($\Phi_{\mathrm{OB}}$\xspace + $\Phi_{\mathrm{HOLMES}}$\xspace), the ionization
parameter $U$, O/H and N/O.
Figure \ref{fig3} shows the location of the observaional points in the
[O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi vs [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi diagram,
with respect to our ``reduced'' grid of models. The dashes lines join
models with same values of log $\Phi_{\mathrm{OB}}$\xspace (equal to 3.5, 4, 4.5, 5, 5.5, 6, 6.5),
while the continuous line join models with the same values of $U$ (equal
to -4, -3.5, -3). Form left to right, each panel corresponds to a
different value of log O/H (-3.9, -3.3, -2.9). The value of N/O is the same in
all the panels, namely -0.5 dex. The observational points that have
the highest values of [O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi and [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi correspond to the largest values of $\Phi_{\mathrm{HOLMES}}$\xspace/$\Phi_{\mathrm{total}}$\xspace.
\begin{figure}
\begin{center}
\includegraphics[width=3.2in]{fig-holmes5.eps}
\caption{``Reduced'' grids (the full grid is 5 time more resolved) of
models in the [O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi vs. [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi plane, for various values of
O/H. See text for details on the varying parameters. The
observational points are superimposed on the grid.}
\label{fig3}
\end{center}
\end{figure}
\section{Results}
For each observational point in the DIG of NGC 891 we select the
models of our finely meshed grid that reproduce simultaneously the
values of [O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi, [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi and [N~{\sc ii}]/H$\alpha$\xspace within the observational uncertainties. We find solutions
for each value of $z$. In Figure \ref{fig4} we show the acceptable
ranges of $\Phi_{\mathrm{HOLMES}}$\xspace/$\Phi_{\mathrm{total}}$\xspace, $U$, $n_\mathrm{e}$
and N/O as a function of $z$. The models clearly indicate a systematic
decrease of the electron density with increasing $z$. If we restrict to
solar metallicity (dark bars in Fig. \ref{fig4}), we find that the models which
fit the observations become dominated by the HOLMES as $z$
increases. Turning the argument around, this might be an indication
that the metallicity of the DIG is roughly solar. Our models also
indicate that N/O increases with increasing destance to the galactic
plane, at least until $|z| \sim$ 1.5 kpc.
\begin{figure}
\begin{center}
\includegraphics[width=4.6in]{fig-holmes7_4.eps}
\caption{$\Phi_{\mathrm{HOLMES}}$\xspace/$\Phi_{\mathrm{OB}}$\xspace, $n_\mathrm{e}$, $U$ and N/O vs $z$ for
the models that fit the observational [O~{\sc iii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi, [O~{\sc ii}]/\ifmmode {\rm H}\beta\xspace \else H$\beta$\xspace\fi, and
[N~{\sc ii}]/H$\alpha$\xspace simultaneously. The green bars indicate the range of the
values for wich the models in the grid fit the observations. The
dark bars show the same, but restricting to solar abundances models.}
\label{fig4}
\end{center}
\end{figure}
\section{Conclusions}
Our scenario, which considers both the population of OB stars and that
of HOLMES, is able to explain the long standing problem of the
ionization of the DIG (see \cite[Flores-Fajardo et al. (2011)]{2011MNRAS.415.2182F}).
|
\section{Introduction} \label{intr}
It is still a challenge for condensed matter community to describe the transition from conventional
Landau Fermi liquid to poorly understood Mott insulator in spite of intense theoretical and experimental studies due to the intricate interplay between metal-insulator transition and ubiquitous magnetic orders. Thus it is rather difficult to treat these two distinct behaviors on an equal footing. Some recent theoretical and experimental works have proposed a possible way to tackle this problem, namely, a possible quantum spin liquid Mott state is obtained at first, then the magnetic ordering is considered to be a secondary effect due to a low-temperature instability of the putative quantum spin liquids. \cite{Senthil1,Senthil2,Vojta,Custers1,Custers2} The proposed quantum spin liquids or its derivatives, the fractional fermi liquids, \cite{Senthil1,Senthil2} do not break any lattice or spin rotation symmetry and may be good candidates for exotic nonmagnetic insulator or metallic states observed in certain organic or heavy fermions compounds. \cite{Powell,Custers1,Custers2,Matsumoto} Therefore, the newly proposal provides a good way to study the nature of Mott metal-insulator transition without extra complexity from broken magnetic long-range ordering.
Generically, the Mott transition from Fermi liquid to a quantum spin liquid is described by slave-particle approaches. \cite{Wen,Florens,Lee,Senthil,Podolsky} In such an approach, the electron is firstly fractionalized into some elementary excitations like spinon, holon, and roton, and so on and then these fractional excitations are bounded with emergent gauge fields to meet the local constraints and to calculate the physical observables.
Recently, an alternative dual description of Mott transition in two space dimensions has been proposed by Mross and Senthil \cite{Mross} based on
the insight from theory of the Mott transition of bosons. In their treatment, low-lying excitations around the two dimensional Fermi surface
are first described by bosonic hydrodynamical modes, then topological defects, namely, the vortices, are identified and inserted into the effective
action of charge phase bosons which are just the hydrodynamical modes being responsible for low energy charge excitations of original Fermi liquid.
If vortices condense, the resulting Mott state is a quantum spin liquid with an emergent gapless $U(1)$ gauge field coupled to a spinons' Fermi
surface while gapped vortices (not condensate) only lead to conventional Landau Fermi liquid.
While their approach only has been applied to systems with two space dimensions, it is interesting to ask a question whether their approach can be extended to higher dimensions, specially physically interesting dimension of three. At a first sight, it seems straightforward to do it. However, a dual description in three space dimensions is rather different from its counterpart in 2D because the point-particle features of topological excitations in 2D are lost in 3D and the conventional boson-vortex duality has to be reformulated with a formidable string field theory instead of a transparent scalar-QED description (a fictitious charged superconductor with a fluctuating $U(1)$ gauge field). Since the description of the charged superconductor is crucial for the conventional boson-vortex duality, it is still unclear whether a reasonable dual description of Fermi liquid exits in the presence of topological defects whose dual description is bosonic string. In the present paper we attempt to answer these questions. We find the dual description of Fermi liquid is still robust. Different to the point-particle feature of the topological excitations in 2D, the topological defects in 3D is line singularities named vortex lines or loops. \cite{Franz,Beekman} Therefore, after identifying the vortex line excitations in charge phase bosons, we obtain an effective description of low-lying Fermi surface fluctuations when refermionizing the corresponding bosonic hydrodynamical action in 3D. The resulting description of three dimensional Mott transition naturally incorporates both conventional Landau Fermi liquid and the expected quantum spin liquid which is formulated in terms of an effective $U(1)$ gauge field minimally coupled to spinons' Fermi surface. Moreover, a doublon metal \cite{Mross} which describes a spin gapped metal with a superconducting instability can also be obtained when the topological defects condense in spin phase degrees of freedom. Interestingly, based on our dual approach, a new state named doubled $U(1)$ gauge theory can occur when vortex lines condense in both phase bosons. This is a consequence of two effective $U(1)$ gauge fields with decoupling of spin-opposite fermions due to their opposite gauge charges.
The remainder of this paper is organized as follows. In Sec.\ref{sec2}, we briefly review basic concepts of boson-vortex duality in four space-time dimensions for bosonic Mott transition and argue the irrelevance of complicated formulism of dual strings description in discussing the duality of three dimensional Fermi liquid in the next section in which we follow the procedure of Mross and Sethil to establish the required hydrodynamical equation for three dimensional Fermi
liquid. Its corresponding bosonized effective action for low-lying excitation of Fermi liquid is derived in Sec.\ref{sec4}. Then the topological vortex
lines excitations are identified in the three dimensional Fermi liquid and incorporated into the charge phase bosons, which leads to an effective description
of Fermi liquid and an anticipated quantum spin liquid in terms of emergent $U(1)$ gauge theory. In addition, we also find a spin gapped metal, a doublon metal, which has an instability to the superconducting pairing below a corresponding critical temperature when topological defects condense in spin phase degree of freedom. Furthermore, when vortex lines condense in both phase bosons, an exotic doubled U(1) gauge theory emerges. Finally, a concise conclusion is devoted to Sec.\ref{sec5}.
\section{Review of boson-vortex duality in four space-time dimensions} \label{sec2}
The conventional boson-vortex duality for bosonic Mott transition in two space dimensions is rather well understood theoretically which links the classic Ginzburg-Landau-Wilson (GLW) paradigm to a fictitious charged superconductor in an emergent fluctuating $U(1)$ gauge field. \cite{Dasgupta,Fisher} The condensed state of the charged superconductor is dual to Mott insulator state of the GLW model which is characterized with the condensation of vortices in the phase degrees of freedom of the effective quantum (2+1)D XY model, a low energy limit of original GLW model. On the other hand, the disordered state of the fictitious superconductor corresponds to the superfluid state of the GLW or quantum XY model and here vortices excitations are gapped with only phonon-like gapless excitations surviving in low energy spectrum.
In the three space dimensions case, the situation is quite different. The vortices form oriented lines or loops, thus the point-particle features of vortices in two dimensional systems are lost and classic boson-vortex duality has to be modified for taking those vortex lines excitations into account. Progress has been made to solve this problem and the resulting dual description involves the celebrated Nambu-Goto, more specifically, the Nielsen-Olesen-type strings which are able to formulate the line singularity of vortices and minimally coupled to the 2-form Kalb-Ramond gauge field. \cite{Franz,Beekman} However, as we will argue here, if we only focus on the original quantum XY model and identify vortex lines in the phase description without immersing into details of duality, it is unnecessary to strive with the formidable string-gauge theory. Instead, we can use the modified quantum XY model-like theory to
proceed our discussion of dual description of three dimensional Fermi liquid-Mott insulator transition.
For our purpose, we begin with the quantum XY model in three space dimensions, which is an effective model of original GLW
action for superfluid-Mott insulator transition with integer-filling:
\begin{equation}
S=\frac{J}{2}\int d\tau d^{3}\vec{x} (\partial_{\mu}\theta(\vec{x},\tau))^{2} \label{eq1}
\end{equation}
where $J$ represents the phase stiffness, $\theta$ is the phase canonically conjugated with local particle density and $\mu = \{\vec{x},\tau\}$.
It is straightforward to use the familiar Hubbard-Stratonovich transformation to decouple the quadratic term in the original quantum XY
model, and Eq. (\ref{eq1}) becomes
\begin{equation}
S=\int d\tau d^{3}\vec{x} \left(\frac{J_{\mu}^{2}}{2J}+iJ_{\mu}\partial_{\mu}\theta\right). \label{eq2}
\end{equation}
If we decompose the phase into a smooth part $\theta_{0}$ and a singular part $\theta_{V}$ which represents the anticipated
vortex lines, we have
\begin{equation}
S=\int d\tau d^{3}\vec{x} \left(\frac{J_{\mu}^{2}}{2J}+iJ_{\mu}(\partial_{\mu}\theta_{0}+\partial_{\mu}\theta_{V})\right). \label{eq3}
\end{equation}
Here we note if one integrates over the auxiliary field $J_{\mu}$, a modified quantum XY model reads
\begin{equation}
S=\frac{J}{2}\int d\tau d^{3}\vec{x} (\partial_{\mu}\theta_{0}+a_{\mu})^{2}. \label{eq4}
\end{equation}
Since the gradient of vortex lines $\partial_{\mu}\theta_{V}$ is a singular function we have replaced it by an effective gauge field $a_{\mu}$.
A cautious reader may wonder why one introduces topological excitations like vortices or vortex lines into quantum XY model above and what role
those topological objects play. The reason is that in the original XY action, the phase variables do not remember they are phase defined mod $2\pi$,
which means the canonically conjugated local density of particles cannot be quantizied into integer per area or volume. Thus, the phase-only description
of original quantum XY action is unable to describe the Mott state. Therefore, in order to have a description of Mott insulator,
one has to impose topological defects such as vortices or vortex lines into the XY model due to their ability to implement the constraint of mod $2\pi$ for
phase variables. Only those topological defects condense, the quantization of density of particles will be fulfilled, which leads to an expected Mott insulator.
Obviously, this modified quantum XY model is general for arbitrary space dimensions because we do not refer to any specific realization
of duality but only make a substitution for incorporating topological vortex lines excitations. Thus, in the discussion of dual approach for
Fermi liquid-Mott insulator transition in the remaining parts of the present paper, it could be safe to follow the same logic as it has been done in this
section without referring specific realization of duality.
For self-contained goal, we would like to state the main results of boson-vortex duality in four space-time dimensions for the condensed matter
community, but eager readers may skip this part without loss of understanding of other sections in this paper.
We continue with the quantum XY action including the auxiliary field $J_{\mu}$, and one can integrate over the smooth phase field $\theta_{0}$, which
leads to a constraint $\partial_{\mu}J_{\mu}=0$. It has been emphasized by Franz \cite{Franz} in three space dimensions that this constraint can only be fulfilled with $J_{\mu}=\epsilon_{\mu\nu\alpha\beta}\partial_{\nu}B_{\alpha\beta}$
where $B_{\mu\nu}$ is a 2-form antisymmetric gauge field and $\epsilon_{\mu\nu\alpha\beta}$ a totally antisymmetric tensor. Therefore, one gets
\begin{equation}
S=\int d\tau d^{3}\vec{x} \left(\frac{H_{\alpha\beta\gamma}^{2}}{3J}+iB_{\alpha\beta}\epsilon_{\mu\nu\alpha\beta}\partial_{\mu}\partial_{\nu}\theta_{V}\right)
\end{equation}
where $H_{\alpha\beta\gamma}=\partial_{\alpha} B_{\beta\gamma}+\partial_{\beta} B_{\gamma\alpha}+\partial_{\gamma} B_{\alpha\beta}$ is a field-strength tensor.
To proceed the dual transformation, the vortex lines degrees of freedom are replaced by the Nielsen-Olsen-type strings \cite{Zwiebach} and their effective action is
\begin{eqnarray}
&& S=\int d\tau d^{3}\vec{x} \frac{H_{\alpha\beta\gamma}^{2}}{3J} + \int \emph{D}[X]\int d\sigma\sqrt{h} \nonumber\\
&&\hspace{0.5cm} \left(\left|\left(\frac{\delta}{\delta\Sigma_{\mu\nu}}
-2\pi B_{\mu\nu}\right)\Phi[X]\right|^{2}+M_{eff}^{2}|\Phi[X]|^{2}\right) \nonumber\\
&& \hspace{0.5cm}+S'_{int},
\end{eqnarray}
where $X$ is the so-called string coordinate, $\sigma$ is the string parameter, $\Phi$ is the string field, $\Sigma_{\mu\nu}$ is the surface element of a worldsheet and $h$ can be considered as an induced metric of the strings
in three space dimensions. Here a mass term $M_{eff}$ and a remaining interacting term $S'_{int}$ which represents short-range string interaction are added.
The physical picture is clear in this situation in spite of the formidable formulism of string-gauge action above. \cite{Franz} The dual description involves the Nielsen-Olesen strings with intrinsic thickness defined by the vortices core sizes and a 2-form
effective gauge field named Kalb-Ramond gauge field in string theory. Therefore, the disorder phase of original GLW or quantum XY action is described in terms of condensation of strings while the original superfluid state corresponds to gapped state of those Nielsen-Olesen strings with free Kalb-Ramond gauge field supporting gapless photonic excitations which in fact play a role as the Goldstone modes
in original symmetry breaking phase of GLW or XY models.
\section{Hydrodynamical description of Fermi liquid} \label{sec3}
Having reviewed the basic of boson-vortex duality, we begin our discussion of the dual description of Fermi liquid-Mott insulator
transition in three space dimensions. Following Mross and Senthil, \cite{Mross} we start with the hydrodynamical equation for conventional Fermi liquid in 3D as follows
\begin{eqnarray}
&& \left(\frac{\partial}{\partial t}+\vec{v}_{\vec{k}}\cdot\frac{\partial}{\partial \vec{x}}\right)\delta n_{\vec{k}\sigma}(\vec{x},t) \nonumber\\
&& \hspace{0.5cm}+\frac{1}{L^{d}}\sum_{\vec{k}'\sigma'}f_{\vec{k}\sigma,\vec{k}'\sigma'}\delta(k_{F}-|\vec{k}|)\hat{\vec{k}}\cdot \frac{\partial}{\partial \vec{x}}\delta n_{\vec{k}'\sigma'}(\vec{x},t)=0,
\end{eqnarray}
where $k_{F}$ is Fermi momentum, $\vec{v}$ is the velocity of quasi-particle with momentum $\vec{k}$, $\delta n_{\vec{k}\sigma}(\vec{x},t)$ is the deviation of ground state distribution function and $f_{\vec{k}\sigma,\vec{k}'\sigma'}$ is the familiar Landau interaction function. Then we introduce local density and total density function in 3D
\begin{equation}
\rho_{\sigma}(\Omega,\vec{x},t)\equiv\int_{-\Lambda}^{+\Lambda}\frac{dk}{(2\pi)^{3}}\delta n_{k,\Omega,\sigma}(\vec{x},t),
\end{equation}
where $\Lambda$ is a cutoff momentum. Here $\delta n_{\vec{k},\sigma}(\vec{x},t)=\delta n_{k,\Omega,\sigma}(\vec{x},t)$, $\Omega\equiv(\theta,\varphi)$, $d\Omega\equiv\int_{0}^{2\pi}d\varphi\int_{0}^{\pi}d\theta\sin\theta$, and
\begin{equation}
\rho_{\sigma}(\vec{x},t)\equiv\int d\Omega k_{F}^{2}(\Omega)\rho_{\sigma}(\Omega,\vec{x},t).
\end{equation}
Consequently, multiplying above hydrodynamical equation with $1/(2\pi)^{3}$ and integrating over $k$, one obtains
\begin{eqnarray}
&& \left(\frac{\partial}{\partial t}+\vec{v}_{F}(\Omega)\cdot\frac{\partial}{\partial \vec{x}}\right)\rho_{\sigma}(\Omega,\vec{x},t) + \frac{1}{(2\pi)^{3}}\sum_{\sigma'}\int d\Omega' k_{F}^{2}(\Omega')\nonumber\\
&& \hspace{1cm}\times f_{\sigma\sigma'}(\Omega,\Omega')\hat{\vec{k}}_{F}(\Omega)\cdot\frac{\partial}{\partial \vec{x}}\rho_{\sigma'}(\Omega',\vec{x},t)=0.
\end{eqnarray}
If we use the notation $\vec{v}_{F}\cdot\frac{\partial}{\partial \vec{x}}=v_{F}\hat{\vec{v}}_{F}\cdot\frac{\partial}{\partial \vec{x}}\equiv v_{F}\partial_{\parallel}$ with $\partial_{\parallel}\equiv\hat{\vec{v}}_{F}(\Omega)\cdot\frac{\partial}{\partial \vec{x}}$ ($\hat{\vec{v}}_{F}$ or $\hat{\vec{k}}_{F}$ is unit vector), we can further simplify above expression into the following form
\begin{eqnarray}
&&\left(\frac{\partial}{\partial t}+v_{F}\partial_{\parallel}\right)\rho_{\sigma}(\Omega,\vec{x},t)\nonumber\\
&& +\frac{1}{(2\pi)^{3}}\sum_{\sigma'}\int d\Omega' k_{F}^{2}(\Omega')f_{\sigma\sigma'}(\Omega,\Omega')\partial_{\parallel} \rho_{\sigma'}(\Omega',\vec{x},t)=0.
\end{eqnarray}
Now we turn into a phase bosons description of Fermi liquid with $\rho_{\sigma}(\Omega,\vec{x},t)=\frac{1}{2\pi}\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t)$, so the equation of motion for Fermi liquid is transformed into
\begin{eqnarray}
&&\left(\frac{\partial}{\partial t}+v_{F}\partial_{\parallel}\right)\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t)\nonumber\\
&& +\frac{1}{(2\pi)^{3}}\sum_{\sigma'}\int d\Omega' k_{F}^{2}(\Omega')f_{\sigma\sigma'}(\Omega,\Omega')\partial_{\parallel}\partial_{\parallel}'\phi_{\sigma}(\Omega',\vec{x},t) = 0.
\end{eqnarray}
\section{The effective bosonized action for Fermi liquid} \label{sec4}
In this section, we proceed to discuss an effective bosonization action for Fermi liquid in 3D which is appropriate for dual description of Mott transition.
It is easy to see that the following action can give rise to the hydrodynamical equation for Fermi liquid in 3D derived in the last section
\begin{equation}
S=\int dt\int d^{3}\vec{x}L(\partial_{t}\phi_{\sigma},\partial_{\parallel}\phi_{\sigma}), \;\; L=L_{0}+L_{I},
\end{equation}
where $L_{0}=\sum_{\sigma}\int d\Omega k^{2}_{F}(\Omega) [\partial_{t}\phi_{\sigma}(\Omega,\vec{x},t) \partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t) - v_{F}(\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t))^{2}]$ and $L_{I} = -\frac{1}{(2\pi)^{3}} \sum_{\sigma\sigma'} \int d\Omega k^{2}_{F}(\Omega)$ $\times \int d\Omega' k^{2}_{F}(\Omega') f_{\sigma\sigma'}(\Omega,\Omega') \partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t) \partial'_{\parallel}\phi_{\sigma'}(\Omega',\vec{x},t)$.
Consider $\phi_{\sigma}(\Omega,\vec{x},t)$ as the canonical coordinate, its corresponding canonical momentum can be written as $\Pi_{\sigma}(\Omega,\vec{x},t)\equiv\frac{\partial L}{\partial(\partial_{t}\phi_{\sigma}(\Omega,\vec{x},t))}=\sin\theta k^{2}_{F}(\Omega)\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},t)$,
therefore we may identify a commutation relation for these two canonically conjugated variables when we turn to the operator framework
\begin{equation}
[\hat{\phi}_{\sigma}(\Omega,\vec{x},t),\hat{\Pi}_{\sigma}(\Omega',\vec{x}',t)]=i\delta(\theta-\theta')\delta(\varphi-\varphi')\delta(\vec{x}-\vec{x}')
\end{equation}
or we have
\begin{eqnarray}
&& [\hat{\phi}_{\sigma}(\Omega,\vec{x},t),2\pi\sin\theta' k^{2}_{F}(\Omega')\hat{\rho}_{\sigma}(\Omega',x,t)] \nonumber\\
&& \hspace{1cm} =i\delta(\theta-\theta')\delta(\varphi-\varphi')\delta(\vec{x}-\vec{x}').
\end{eqnarray}
It is straightforward to derive a further commutation relation if one defines the local particle density operator ($\hat{\rho}_{\sigma}(\vec{x},t)\equiv\int d\Omega k_{F}^{2}(\Omega)\hat{\rho}_{\sigma}(\Omega,\vec{x},t)$)
and the local phase operator ($\hat{\phi}_{\sigma}(\vec{x},t)=\frac{1}{4\pi}\int d\Omega \hat{\phi}_{\sigma}(\Omega,\vec{x},t)$)
\begin{equation}
[\hat{\phi}_{\sigma}(\vec{x},t),2\pi \hat{\rho}_{\sigma}(\vec{x}',t)]=i\delta(\vec{x}-\vec{x}').
\end{equation}
This commutation relation will be useful in next section where we introduce topological objects named vortex lines.
Having the real-time action in hand, we can also transform it into imaginary-time action which reads as follows
\begin{equation}
S=\int d\tau\int d^{3}\vec{x}L(\partial_{\tau}\phi_{\sigma},\partial_{\parallel}\phi_{\sigma}), \;\; L=L_{0}+L_{I}
\end{equation}
where $ L_{0}=\sum_{\sigma}\int d\Omega k^{2}_{F}(\Omega)
[-i\partial_{\tau}\phi_{\sigma}(\Omega,\vec{x},\tau)\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},\tau)$ $+ v_{F}(\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},\tau))^{2}]$ and $ L_{I}=\frac{1}{(2\pi)^{3}}\sum_{\sigma\sigma'}\int d\Omega\int d\Omega' $ $k^{2}_{F}(\Omega) k^{2}_{F}(\Omega')f_{\sigma\sigma'}(\Omega,\Omega')\partial_{\parallel}\phi_{\sigma}(\Omega,\vec{x},\tau)\partial'_{\parallel}\phi_{\sigma'}(\Omega',\vec{x},\tau)$.
As done in the situation of the hydrodynamical description for one dimensional fermions systems (i.e. the Luttinger liquid), \cite{Gogolin} one can define charge and spin phase variables (bosons)
\begin{eqnarray}
&&\phi_{c}(\Omega,\vec{x},\tau)=\frac{1}{2}(\phi_{\uparrow}(\Omega,\vec{x},\tau)+\phi_{\downarrow}(\Omega,\vec{x},\tau)), \\
&&\phi_{s}(\Omega,\vec{x},\tau)=\frac{1}{2}(\phi_{\uparrow}(\Omega,\vec{x},\tau)-\phi_{\downarrow}(\Omega,\vec{x},\tau)).
\end{eqnarray}
Correspondingly, one can also introduce charge and spin density variables
\begin{eqnarray}
&&\rho_{c}(\Omega,\vec{x},\tau)=\frac{1}{2}(\rho_{\uparrow}(\Omega,\vec{x},\tau)+\rho_{\downarrow}(\Omega,\vec{x},\tau)), \\
&&\rho_{s}(\Omega,\vec{x},\tau)=\frac{1}{2}(\rho_{\uparrow}(\Omega,\vec{x},\tau)-\rho_{\downarrow}(\Omega,\vec{x},\tau)),
\end{eqnarray}
where $\rho_{c}=\frac{1}{2\pi}\partial_{\parallel}\phi_{c}$ and $\rho_{s}=\frac{1}{2\pi}\partial_{\parallel}\phi_{s}$.
Using above charge and spin phase variables, the effective action for Fermi liquid in 3D can be regrouped into three terms which represent charge, spin excitations, and their interactions, respectively,
\begin{eqnarray}
&& S=S_{c}+S_{s}+S_{c-s}, \\
&& S_{c}=\int d\tau\int d^{3}\vec{x}(L_{c}^{0}+L_{c}^{I}), \\
&& S_{s}=\int d\tau\int d^{3}\vec{x}(L_{s}^{0}+L_{s}^{I}), \\
&& S_{c-s}=\frac{1}{4\pi^{3}}\int d\Omega\int d\Omega' k^{2}_{F}(\Omega) k^{2}_{F}(\Omega')\nonumber\\
&& \hspace{1cm} \times (f_{\uparrow\uparrow}-f_{\downarrow\downarrow})\partial_{\parallel}\phi_{c}(\Omega,\vec{x},\tau)\partial'_{\parallel}\phi_{s}(\Omega',\vec{x},\tau).
\end{eqnarray}
Here the notations $L_{c,s}^{0}, L_{c,s}^{I}$ are introduced as
$ L_{c}^{0}=\int d\Omega k^{2}_{F}(\Omega) [-i\partial_{\tau}\phi_{c}(\Omega,\vec{x},\tau)\partial_{\parallel}\phi_{c}(\Omega,\vec{x},\tau)+v_{F}(\partial_{\parallel}\phi_{c}(\Omega,\vec{x},\tau))^{2}]$, $L_{c}^{I}=\frac{1}{4\pi^{3}}\int d\Omega\int d\Omega' k^{2}_{F}(\Omega) k^{2}_{F}(\Omega')$ $f_{s}(\Omega,\Omega') \partial_{\parallel}\phi_{c}(\Omega,\vec{x},\tau)\partial'_{\parallel}\phi_{c}(\Omega',\vec{x},\tau)$, $L_{s}^{0}=\int d\Omega k^{2}_{F}(\Omega)
$ $[-i\partial_{\tau}\phi_{s}(\Omega,\vec{x},\tau)\partial_{\parallel}\phi_{s}(\Omega,\vec{x},\tau)+v_{F}(\partial_{\parallel}\phi_{s}(\Omega,\vec{x},\tau))^{2}]$, $L_{s}^{I}=\frac{1}{4\pi^{3}}\int d\Omega\int d\Omega' k^{2}_{F}(\Omega) k^{2}_{F}(\Omega')f_{a}(\Omega,\Omega') \partial_{\parallel}\phi_{s}(\Omega,\vec{x},\tau) $ $\partial'_{\parallel}\phi_{s}(\Omega',\vec{x},\tau)$,
and $f_{s}=(f_{\uparrow\uparrow}+f_{\downarrow\uparrow})/2$ and $f_{a}=(f_{\uparrow\uparrow}-f_{\uparrow\downarrow})/2$ are standard Landau interaction function, respectively. In particular, if we further assume a paramagnetic solution of Fermi liquid, the charge-spin interaction term will vanish in this situation and we will obtain an effective action with seemingly spin-charge separation formulism.
In the remaining part of the present paper, we only consider this paramagnetic Fermi liquid for simplicity.
\section{Incorporating vortex lines into phase bosons} \label{sec5}
It is well-known that vortices are favorable in classical or quantum XY model-like systems in space dimension of two when thermal or
quantum fluctuation are strong enough to suppress ordered states. If these vortices proliferate, they will drive the whole system into thermal
or quantum disordered phase and long-range order will disappear. Particularly, if the original bosonic Hamiltonian filled with integral number of particles,
the vortices condensation induced transition should be superfluid-Mott transition with Mott state identified by condensate of vortices. However,
in 3D point-particle-like vortices are not energetically preferable while topological vortex lines excitation may drive the
Mott transition of bosons. We have briefly reviewed the basis of the dual approach of this Mott transition in the previous section and based on those
insights we will follow the same logic of Mross and Senthil \cite{Mross} to fulfill
our dual description of Fermi liquid-Mott insulator transition in space dimension of three in this section.
As mentioned above, Fermi liquid can be rewritten in terms of bosonized effective action which only
incorporates low energy charge and spin excitation around Fermi surface. Based on insights from the above discussion of vortex lines driven superfluid-Mott transition of bosons and stimulated by work of Mross and Senthil for two dimensional Fermi liquid-Mott insulator transition, we also
expect that a similar Fermi liquid-Mott insulator transition could be driven by condensation of vortex lines in space dimension of three. Now we first proceed to identify vortex lines in charge phase bosons $\phi_{c}$. As emphasized by Mross and Senthil, \cite{Mross} the correct procedure of incorporating topological object is to firstly identify local charge phase which is canonically conjugated to local charge density, then to insert topological objects such as vortices or vortex lines into local charge phase. It is easy to see the local charge phase is a simple addition of local phase for different spins
\begin{equation}
\phi_{c}(\vec{x},\tau)=\frac{1}{2}(\phi_{\uparrow}(\vec{x},\tau)+\phi_{\downarrow}(\vec{x},\tau))=\frac{1}{4\pi}\int d\Omega \phi_{c}(\Omega,\vec{x},t).
\end{equation}
Here it is useful to expand $\phi_{c}(\Omega,\vec{x},t)$ in terms of spherical harmonic functions $Y_{lm}(\Omega)$
\begin{equation}
\phi_{c}(\Omega)=\sum_{lm}Y_{lm}(\Omega)\phi_{c,lm},
\end{equation}
where $\phi_{c,lm}=\frac{1}{4\pi}\int d\Omega Y^{\ast}_{lm}(\Omega)\phi_{c}(\Omega)$. We note that $\phi_{c,l=m=0}=\frac{1}{4\pi}\int d\Omega \phi_{c}(\Omega)$ is just the local charge phase we need. Therefore, following Mross and Senthil, we need to separate the mode with $l=m=0$ from other modes in the bosonic effective action. After a straightforward manipulation, we obtain the following action
\begin{equation}
S_{c}=\int d\tau\int d^{3}\vec{x}L_{c},
\end{equation}
where
\begin{widetext}
\begin{equation}
L_{c}=-i\sum_{lml'm'\neq0}\partial_{\tau}\phi_{c,lm}\vec{Z}_{mm'}^{ll'}\partial_{\vec{x}}\phi_{c,l'm'}-2i\sum_{lm\neq0}\partial_{\tau}\phi_{c,00}\vec{Z}_{m0}^{l0}\partial_{\vec{x}}\phi_{c,lm}
+\sum_{lm,l'm'}\partial_{\alpha}\phi_{c,lm} M_{lm,l'm'}^{\alpha\beta}\partial_{\beta}\phi_{c,l'm'}.
\end{equation}
Here $\vec{Z}_{mm'}^{ll'}=\int d\Omega k_{F}^{2}(\Omega)Y_{lm}(\Omega)Y_{l'm'}(\Omega)\hat{\vec{v}}_{F}(\Omega)$ and $M_{lm,l'm'}^{\alpha\beta}=2v_{F}\int d\Omega k_{F}^{2}(\Omega)\hat{v}_{F}^{\alpha}(\Omega)\hat{v}_{F}^{\beta}(\Omega)Y_{lm}(\Omega)Y_{l'm'}(\Omega)+\frac{1}{4\pi^{3}}\int d\Omega\int d\Omega'k_{F}^{2}(\Omega)k_{F}^{2}(\Omega')f_{s}(\Omega,\Omega')\hat{v}_{F}^{\alpha}(\Omega)\hat{v}_{F}^{\beta}(\Omega')Y_{lm}(\Omega)Y_{l'm'}(\Omega')$.
Consequently, we insert the vortex lines degrees of freedom into the mode of $l=m=0$ with the direct substitution $\phi_{c,lm}\rightarrow\phi_{c,lm}+\phi_{V}\delta_{l=m=0}$. Since the vortex lines are singular manifolds, we replace its gradient $\partial_{\mu}\phi_{V}$ with an effective $U(1)$ gauge field $a_{\mu}$.
Therefore an effective action for charge degrees of freedom of Fermi liquid can be derived as
\end{widetext}
\begin{equation}
S_{c}=\int d\tau\int d^{3}\vec{x} \int d\Omega k_{F}^{2} (L\{\partial_{\tau}\phi_{c},\partial_{\parallel}\phi_{c}\}+j_{\mu}a_{\mu})
\end{equation}
and for spin section, the effective action can be written as
\begin{equation}
S_{s}=\int d\tau\int d^{3}\vec{x} \int d\Omega k_{F}^{2} (L\{\partial_{\tau}\phi_{s},\partial_{\parallel}\phi_{s}\}).
\end{equation}
Combining above two parts for the whole effective action of Fermi liquid, we have
\begin{equation}
S=\int d\tau\int d^{3}\vec{x} \int d\Omega k_{F}^{2} (L\{\partial_{\tau}\phi_{\sigma},\partial_{\parallel}\phi_{\sigma}\}+ j_{\mu}^{\sigma}a_{\mu}).
\end{equation}
It is well know that the action without vortex lines describes conventional Fermi liquid if one refermionizes the bosonic
action by using bosonization identity $f_{\sigma}(\vec{x},\tau,\Omega)\sim e^{i(2\pi)^{3/2}\phi_{\sigma}(\vec{x},\tau,\Omega)}$ with $f_{\sigma}$ representing a fermion with a spin. As a matter of fact, we find a more general bosonization relation for refermionizing phase bosons of original Fermi liquid situated in arbitrary dimension $d$ with its form being $f_{\sigma}(\vec{x},\tau,\Omega)\sim e^{i(2\pi)^{d/2}\phi_{\sigma}(\vec{x},\tau,\Omega)}$. In the low energy limit, all interaction terms in bosonic action between bosons of two patches near Fermi surface is irrelevant and
we may use free fermions' Lagrangian for the term without vortex lines
\begin{equation}
L\{\partial_{\tau}\phi_{\sigma},\partial_{\parallel}\phi_{\sigma}\}=f_{\sigma}^{+}\left(\partial_{\tau}-\mu_{F}+\frac{-\nabla^{2}}{2m}\right)f_{\sigma}
\end{equation}
and the whole Lagrangian can be rewritten as
\begin{equation}
L=f_{\sigma}^{+}\left(\partial_{\tau}-\mu_{F}-ia_{0}+\frac{-(\nabla-i\vec{a})^{2}}{2m}\right)f_{\sigma},
\end{equation}
where $\mu_{F}$ and $m$ are effective chemical potential and mass of the fermions, respectively. We notice that the above action is just the one obtained in continuous Mott transition for generic systems described by Hubbard model
nearly half-filling \cite{Podolsky} and Kondo-breakdown mechanism \cite{Senthil2,Vojta,Paul,Pepin} for heavy fermions compounds like $YbRh_{2}Si_{2}$ and $YbRh_{2}Si_{2-x}Ge_{x}$ \cite{Custers1,Custers2} in 3D with the help of slave rotor or slave boson representations. This action describes fermion interacting with each other in terms of an effective $U(1)$ gauge field
in the background of a Fermi surface of those fermions. In fact, these fermions are fermionic spinons without carrying charges due to
condensation of vortex lines in charge phase bosons. The condensation gaps the charge excitation in low energy and the whole theory describe
a $U(1)$ quantum spin liquid with a spinons' Fermi surface. This picture has been proposed to describe the nonmagnetic state down to lowest temperature
in organic Mott insulator. \cite{Lee,Powell}
It should be emphasized that the above $U(1)$ gauge theory action implicitly assumes a deconfined state for the $U(1)$ gauge field which is believed to be possible in 3D. This is in contrast to the situation of 2D where de-confinement is only faithfully established near
some bosonic quantum critical point besides various artificial large-N arguments in the presence of Fermi surface. \cite{Kogut,Polyakov1,Polyakov2,Senthil3,Senthil4,Kim,Lee2} Otherwise, if $U(1)$ gauge field is confined,
the above action is actually meaningless because only bound states of fermions are able to appear in low energy spectrum due to the linear confined potential mediated by confined gauge field between a pair of fermions.
Therefore, the insulating state with confined gauge field cannot be described by a quantum spin liquid but may be a valence bond solid or more conventional spin ordered phases like collinear anti-ferromagnetic states. \cite{Sachdev1,Sachdev2}
Moreover, a doublon metal obtained by Mross and Senthil \cite{Mross} with a spin gap can also be found when vortex lines condense in spin but not in charge phase bosons in three space dimensions. This metal state is described by the Lagrangian
\begin{equation}
L=\sum_{\sigma=\pm1} f_{\sigma}^{+}\left(\partial_{\tau}-\mu_{F}-i\sigma a_{0}+\frac{-(\nabla-i\sigma\vec{a})^{2}}{2m}\right)f_{\sigma},
\end{equation}
where fermions $f_{\sigma}$ with distinct spins couple to a gauge field $b_{\mu}$ with opposite gauge charges, thus gives rise to an attractive interaction between opposite charged fermions. The doublon metal mentioned above is originally derived from a fluctuating doped anti-ferromagnet, which has been applied to the cuprate superconductors. \cite{Kaul,Galitski,Moon} Although a paring instability will be induced by the mediated attractive interaction in this model, we could consider this metal state as a nontrivial intermediate temperature state above the superconducting critical temperature, which is a non-Fermi liquid state with anomalous thermal and transport behaviors compared to conventional Landau Fermi liquid.
Up to now, we have found that three distinct states by our dual approach, they are the conventional Fermi liquid, a $U(1)$ quantum spin liquid and a doublon metal, respectively. Since vortex lines could condense in charge or spin phase bosons, it may be reasonable to expect a double condensation in both phase bosons. It is straightforward to find a doubled gauge theory, which reads as follows:
\begin{eqnarray}
&& L=f_{\uparrow}^{+}\left(\partial_{\tau}-\mu_{F}-i(a_{0}+b_{0})+\frac{-(\nabla-i(\vec{a}+\vec{b}))^{2}}{2m}\right)f_{\uparrow} \nonumber \\
&& +f_{\downarrow}^{+}\left(\partial_{\tau}-\mu_{F}-i(a_{0}-b_{0})+\frac{-(\nabla-i(\vec{a}-\vec{b}))^{2}}{2m}\right)f_{\downarrow}.
\end{eqnarray}
In this Lagrangian, fermions with different spins have the same gauge charges when couple to gauge field $a_{\mu}$ but carry opposite gauge charges with respect to gauge field $b_{\mu}$. Since the obtained Lagrangian is a new result of this paper, here we would like to analyze its physical properties briefly. Firstly, at mean field level where one simply neglects those two gauge fields, an expected gas of free fermions is reproduced. Then, when one turns up the coupling between fermions and two kinds of gauge fields, interestingly, the effective interaction vanishes between fermions with opposite spins mediated by gauge field $a_{\mu}$ and $b_{\mu}$, due to difference in the sign of gauge charges of fermions at the one loop level. Thus, fermions with opposite spins are approximately decoupled to the leading order of perturbation theory and the remaining parts of the effective Lagrangian can be treated separately as done in the Kondo-breakdown mechanism \cite{Senthil2,Vojta,Paul,Pepin} of heavy fermions systems. To our knowledge, this effective action has not been realized by any realistic lattice models, but we hope this exotic quantum state may be useful for further study.
It is noted recently, the derivative of strings theory, the anti-de Sitter/conformal field theory (AdS/CFT) correspondence, has been used to attack the perplexing non-Fermi liquid behaviors in strongly correlated electronic systems \cite{Faulkner1,Faulkner2,Liu} and a clear correspondence is found between certain quantum gravity theories with an emergent AdS$_{2}\times R^{d-1}$ geometry and fractionalized metallic states of Kondo-Heisenberg model, \cite{Sachdev3} thus we hope similar physics as established in terms of our dual approach above may be inspected through this promising approach.
\section{Conclusion}
In conclusion, we have extended the dual approach of two dimensional Fermi liquid-Mott insulator transition
into the physically interesting space dimension of three in spite of complicated details of the dual description with
vortex lines involved. A quantum spin liquid state has been found in three space dimensions when vortex lines condense
in charge phase bosons while the gapped vortex lines only lead to conventional Landau Fermi liquid. In addition, we also have found a spin gapped metal state called as doublon metal, which may support an unconventional superconductivity when topological defects condense in spin phase degrees of freedom. In addition, a double condensation of vortex lines in both phase bosons has been examined and an exotic doubled $U(1)$ gauge theory emerges in this case. This result describes a decoupling of spin-opposite fermions due to their opposite gauge charges. Therefore, four distinct states have been found in our dual approach although details of these states and their corresponding subtle quantum phase transitions have not been captured by this method, which deserves further study in the future. We hope this alternative dual description may provide more insights into Fermi liquid-Mott insulator transition in space dimension of three. Furthermore, similar physics as established in terms of our dual approach above could be inspected through the promising AdS/CFT correspondence since the holographic realization of the familiar compressible quantum states of condensed matter physics, superfluid and Fermi liquid, have been found firmly. \cite{Sachdev4}
\begin{acknowledgments}
The authors would like to thank Y. Q. Wang for useful discussion on Nambu-Goto strings and related issues of duality. The work is supported partly by the Program for NCET, NSF and the Fundamental Research Funds for the Central Universities of China.
\end{acknowledgments}
|
\section{Thermodynamics as a resource theory}
\label{sec:thermo_resource}
In the micro-regime, when the amount of work which can be extracted might be of the order of $kT$, we need to
very precisely define what we mean by work, and what processes are allowed during the extraction of work from
a system. For our purposes, obtaining work $\Delta W$ means to obtain an eigenstate of the Hamiltonian with energy $W_{out}$
starting from an eigenstate of energy $W_{in}$, where $W_{in}-W_{out}=\Delta W$.
In our approach it will turn out, that the amount of work we can extract from a given system
does not depend on the Hamiltonian of the system which stores the work, and the particular levels we choose.
We can thus consider a system of the smallest dimension, which carries work $W$.
This is a two level system with Hamiltonian $\hat{W}=W |1\>\<1|$. We shall call this a {\it work qubit} (in short, a {\it wit)},
and let $\ket{\psi_W}$ denote the excited state $\ket{1}$ with energy $W$.
This is the most economical way of storing work.
Since drawing or adding work can be represented as a state transformation, it is natural to consider
thermodynamics as a resource theory. Namely, one considers some class of operations,
and then asks how much of some resource can be obtained. Recent examples of such theories include entanglement theory\cite{BBPS1996,thermo-ent2002},
thermodynamics with no Hamiltonian\cite{uniqueinfo}, thermodynamics of erasure\cite{janzing_quasi-order_2003}
and operations which respect a symmetry\cite{gour_resource_2008,marvian_pure_2011}.
Here, we use the class of operations which
corresponds to thermodynamics\cite{janzing_quasi-order_2003,thermoiid}, and then ask by how much we can excite a system
initially in a pure ground state.
It can be shown that there are a number of equivalent ways
of describing this class of operations\cite{thermoiid}.
Since we are interested in extracting work in the presence of a heat bath, one starts by allowing
a free resource of a heat bath, with Hilbert space ${\cal H}_R$. The heat bath is in a
Gibbs state $\tau$, with arbitrary Hamiltonian
and we further allow the addition of any auxiliary system $S'$ with Hamiltonian $H_{S'}$
in a Gibbs state.
Without loss of generality, we can take the initial Hamiltonian to be non-interacting at very early times between the reservoir $R$ and
the system of interest $S$, as well as any ancillas. We also want that initially (and finally), the work qubit
is not interacting with the rest of the system, since we want to be able to store the work, and use it in some other process.
We thus have initially $H_{tot}=H_R+ H_S +H_{S'} + \hat{W}$.
We now require that all manipulations conserve energy. This ensures that all sources of work are properly accounted for, and that
external systems are not adding or taking away work.
The dynamics can be implemented by an interaction Hamiltonian, however, if we wish to maintain a precise accounting of all energy,
then the interaction term needs to vanish at the beginning and end of the protocol, otherwise it allows us to pump work into the system at no cost.
Essentially we need to ensure conservation of total energy. This also means that if we wish to model
a time-dependent Hamiltonian, we should do so by means of a time-independent Hamiltonian with a clock included in the system. It is not
difficult to show \cite{thermoiid}, that all of these paradigms which conserve energy, are equivalent to unitary transformation
commuting with the total Hamiltonian. Essentially, since accounting for all sources of energy requires that the initial and final
Hamiltonian are the same, the dynamics must map eigenstates of the Hamiltonian to eigenstates with the same energy. This is equivalent to
considering a fixed Hamiltonian, and allowing operations which commute with the Hamiltonian.
We also allow discarding subsystems (partial trace). We call this class - {\it Thermal Operations}.
Note that this paradigm allows one to include time independent Hamiltonians as in the example discussed in the Main Section
\begin{align}
H_{tot}=\proj{0}_C\otimes H + \proj{1}_C\otimes H' + W\proj{1}
\label{eq:changingham2}
\end{align}
Via a similar mechanism, one can include interacting terms which vanish at early and late times.
Generally, we are interested in transitions between $(\rho_S, H_S)$ and $(\sigma_{S'}$ and $H_{S'})$
(extracting work will be a special case of such a transition). Since in the described approach,
the Hamiltonian is fixed, such a transition means actually
$(\rho_S\otimes\tau_{S'}, H_S+H_{S'})\to (\tau_{S} \otimes \sigma_{S'}, H_S+H_{S'})$.
where we have the same initial and final Hamiltonian.
\subsection{Assumptions on heat bath, and its relation to the system}
We also assume that Hamiltonians of all systems of concern
(i.e. heat bath Hamiltonian, auxiliary systems, the resource system itself) have minimal energy zero.
Let $E_R$ be energies of reservoir, and $E_S$ be energies of the system.
Let $E_R^{\max}$, and $E_S^{\max}$ be the largest energy of the heat bath and system, respectively
(of course a typical heat bath will have $E_R^{\max}=\infty$).
Our heat bath will be large, while our resource states will be small.
This means that the system Hilbert space will be fixed, while the
energy of the heat bath (and other relevant quantities such as size of degeneracies) will tend to infinity.
We now make some assumptions concerning the state and Hamiltonian of the heat bath.
The heat bath is in a Gibbs state with inverse temperature $\beta$. Moreover there exists
set of energies ${\ecal_R}$ such that
the state of the heat bath occupies energies from ${\ecal_R}$ with high probability, i.e.
for the projector $P_{\ecal_R}$ onto the states with energies ${\ecal_R}$ we have
\begin{eqnarray} \begin{aligned}
\mathop{\mathsf{tr}}\nolimits P_{\ecal_R} \rho_R\geq 1-\delta
\end{aligned} \end{eqnarray}
and it has the following properties:
\begin{enumerate}[(i)]
\item The energies $E$ in ${\ecal_R}$ are peaked around some mean value, i.e.
they satisfy $E\in \{\<E\>- O(\sqrt{\<E\>}),\ldots \<E\>+O(\sqrt{\<E\>})\}$
\label{item:peaked}
\item For $E\in{\ecal_R}$ the degeneracies $g_R(E)$ scale exponentially with $E$, i.e.
\begin{eqnarray} \begin{aligned}
g_R(E) \geq e^{c E}
\end{aligned} \end{eqnarray}
where $c$ is a constant.
\label{item:g_exp}
\item For any three energies $E_R,E_S$ and $E_s'$ such that $E_R\in{\ecal_R}$ and
$E_S$, $E_S'$ are arbitrary energies of the system, there exist $E'_R \in{\ecal_R}$
such that $E_R +E_S = E_R'+E_S'$.
\label{item:matching}
\item For $E\in{\ecal_R}$ the degeneracies $g_R(E)$ satisfy $g_R(E-E_S) \approx g_R(E) e^{-\beta E_S}$,
or more precisely:
\begin{eqnarray} \begin{aligned}
\left|\frac{g_R(E)e^{-\beta E_S'}}{g_R(E-E_S)}-1\right|\leq \delta
\label{eq:g-property}
\end{aligned} \end{eqnarray}
for all energies $E_S$ of the system $S$.
\label{item:g-property}
\end{enumerate}
{\it Discussion of assumptions:}
\begin{itemize}
\item[Ad. \eqref{item:peaked}] This is a standard property of a heat bath.
\item[Ad. \eqref{item:g_exp}] Follows from the condition \eqref{item:peaked} of small fluctuations
combined with extensivity of energy.
\item[Ad. \eqref{item:matching}] Follows from continuity of the spectrum of the heat bath, which is usually the case.
\item[Ad. \eqref{item:g-property}] Follows from
\begin{align}
g(E+\Delta E)&=e^{S(E+\Delta E)}\nonumber\\
&\approx e^{S(E)+\Delta E \frac{\partial S(E)}{\partial E}}\nonumber\\
&=g(E) e^{\beta \Delta E}
\end{align}
with $S(E):=\ln g(E)$. and $\beta:=\frac{\partial S(E)}{\partial E}$.
\end{itemize}
It is also easy to see that a product $\tau^{\otimes n}$ of many copies of independent
Gibbs states satisfies the above assumptions.
\section{Notation and preliminary facts.}
\label{sec:prelim}
We shall now need a bit of notation. Let us define $\eta^X_E$ as a state of a system X proportional to
the projection on to a subspace of energy $E$ (according to the Hamiltonian $H_X$ on this system).
In particular, $\eta_{E-E_S}$ is given by
\begin{eqnarray} \begin{aligned}
\eta_{E-E_S}=g(E-E_S)^{-1}\sum_g |E-E_S,g\>_R\<E-E_S,g|
\end{aligned} \end{eqnarray}
where $g=1,..,g(E-E_S)$, i.e. $\eta_{E-E_S}$ is the maximally mixed state of the reservoir with
support on the subspace of energy $E_R=E-E_S$.
We shall also use notation $\eta_K=\mathbb{I}/ K$ where the identity acts on a $K$ dimensional space.
Let us note that the total space ${\cal H}_R\otimes {\cal H}_S$ can be decomposed as follows
\begin{eqnarray} \begin{aligned}
\label{eq:Hoplus}
{\cal H}_R\otimes {\cal H}_S= \bigoplus_E \left(\bigoplus_{E_S} {\cal H}^R_{E-E_S} \otimes {\cal H}^S_{E_S} \right)
\end{aligned} \end{eqnarray}
(here for $E\leq E_S$ and $E\geq E_R^{\max} +E_S^{\max}$ the summation over $E_S$
is suitably constrained, however we are interested only in energies $E_R$ from ${\ecal_R}$,
hence these cases will not occur).
Consider an arbitrary state $\rho_{RS}$ which has
support within $E_S^{\max}\leq E\leq E_R^{\max}$. We can rewrite it as follows
\begin{eqnarray} \begin{aligned}
\rho_{RS}=\sum_E \sum_\Delta P_E \rho_{RS} P_{E+\Delta}\\
\end{aligned} \end{eqnarray}
Here $\Delta= -E_S^{\max}, \ldots , E_S^{\max}$. The blocks
$P_E \rho_{RS} P_{E+\Delta}$ we can further divide into sub-blocks
\begin{eqnarray} \begin{aligned}
P_E \rho_{RS} P_{E+\Delta}=
\sum_{E_S\in I_\Delta} \mathbb{I}_R \otimes P_{E_S} P_E \rho_{RS} P_{E+\Delta} \mathbb{I}_R \otimes P_{E_S+\Delta}
\end{aligned} \end{eqnarray}
where $I_\Delta=\{0,\ldots, E_S^{\max}-\Delta\}$ for $\Delta \geq 0$ and
$I_\Delta=\{-\Delta,\ldots, E_S^{\max}\}$ for $\Delta \leq 0$.
The sub-blocks map the Hilbert space ${\cal H}^R_{E-E_S}\otimes ({\cal H}^S_{E_S+\Delta}$ onto
${\cal H}^R_{E-E_S}\otimes {\cal H}^S_{E_S}) $
We can then extract the state $\rho_S$
\begin{eqnarray} \begin{aligned}
\rho_S=\sum_{E_S,E_S'} P_{E_S} \rho_S P_{E_S'}
\end{aligned} \end{eqnarray}
as follows:
\begin{eqnarray} \begin{aligned}
P_{E_S} \rho_S P_{E_S'} = \sum_{E} \mathop{\mathsf{tr}}\nolimits_{{\cal H}^R_{E-E_S}}
(P^R_{E-E_S} \otimes P_{E_S} P_E \rho_{RS} P_{E+E_S'-E_S} P^R_{E-E_S} \otimes P_{E_S'})
\end{aligned} \end{eqnarray}
We then have the following technical result that will be a basis for most of our derivations:
\begin{theorem}
\label{thm:oplus_E}
We consider set of energies
\begin{eqnarray} \begin{aligned}
{\cal E}=\{E:E-E_S\in {\ecal_R}\}
\end{aligned} \end{eqnarray}
where ${\ecal_R}$ satisfies assumptions \eqref{item:peaked}, \eqref{item:g_exp} and \eqref{item:matching} listed above. Then
\begin{eqnarray} \begin{aligned}
\forall{E\in {\cal E}} \quad
||\frac{1}{p_E}P_E \rho_R\otimes\rho_S P_{E+\Delta} - \oplus_{E_S}\eta_{E-E_S}\otimes P_{E_S} \rho_S P_{E_S}||\leq 2\delta
\label{eq:oplus_E}
\end{aligned} \end{eqnarray}
and
\begin{eqnarray} \begin{aligned}
\sum_{E\in{\cal E}} p_E\geq 1 -2\delta
\end{aligned} \end{eqnarray}
where $p_E=\mathop{\mathsf{tr}}\nolimits(P_E \rho_R \otimes \rho_S)$.
\end{theorem}
{\it Proof.} Here we sketch the proof for $\Delta=0$. For $\Delta\not=0$
the proof is similar.
Let us fix an energy block $E$.
Let $E_R=E-E_S$. The state $\tau_R \otimes \rho_S$ restricted to
the energy $E$ block is given by
\begin{eqnarray} \begin{aligned}
P_E \tau_R \otimes \rho_S P_E=\frac{1}{Z_R}\sum_{E_S} e^-{\beta E-E_S } g_R(E-E_S) \mathbb{I}_R^{E-E_S} \otimes P_{E_S} \rho_S P_{E_S}
\end{aligned} \end{eqnarray}
where $Z$ is partition function for system $R$, and $\mathbb{I}_R^{E-E_S}$ is identity on the subspace ${\cal H}_R^{E-E_S}$, see \eqref{eq:Hoplus}.
Using \eqref{item:g-property} we have $g_R(E-E_S)=g_R(E) e^{-\beta E_S}$ we get
\begin{eqnarray} \begin{aligned}
P_E \tau_R \otimes \rho_S P_E \approx \frac{1}{Z_R} e^{-\beta E} g_R(E)
\sum_{E_S} \frac{\mathbb{I}_R^{E-E_S}}{g_R(E)e^-{\beta E_S}}\otimes P_{E_S} \rho_S P_{E_S}
\end{aligned} \end{eqnarray}
Since $\frac{\mathbb{I}_R^{E-E_S}}{g_R(E)e^-{\beta E_S}}=\eta_{E-E_S}$. Moreover, if we drop the prefactor,
the state is normalised, hence we obtain the claim.
\section{Transformations of classical states: condition in terms of majorization}
\label{sec:major}
Here we will provide a necessary and sufficient condition for transforming the
diagonal part of a density matrix of one state into the diagonal part of another state acting on the same system.
The condition will be in terms of the so called {\it majorization} condition, and it will be necessary and sufficient for
state transformations of classical states (i.e. diagonal in the energy eigenbasis).
The result is contained in the theorem \ref{thm:major}.
From the expression \eqref{eq:oplus_E} it follows that a block of fixed energy $E$
contains only the diagonal part of $\rho_S$:
\begin{eqnarray} \begin{aligned}
P_E \rho_R\otimes \rho_S P_E \approx \oplus_{E_S}\eta_{E-E_S}\otimes P_{E_S} \rho_S P_{E_S}
\label{eq:eta_form}
\end{aligned} \end{eqnarray}
Note that we can tensor out a maximally mixed state of size independent of both $E_S$ and $E$,
and apply unitaries conditioned on the maximally mixed state. We do that by
writing
\begin{eqnarray} \begin{aligned}
\eta_{E-E_S}=\frac{\mathbb{I}_K}{K} \otimes \frac{\mathbb{I}_{K'_{E_S}}}{K'_{E_S}}
\end{aligned} \end{eqnarray}
where $K'_{E_S} = g(E-E_S)/K$.
Due to our assumptions about the heat bath, the degeneracy of each energy state is exponentially large in energy,
so we can take such $K$ that both $K$ and $K'$ are exponentially large in energy.
Thus a given fixed energy block $E$ can be represented as a tensor product of two systems
$R_1^E$ and $R_2^E S^E$ in a state
\begin{eqnarray} \begin{aligned}
\frac{1}{p_E} P_E \rho_R\otimes \rho_S P_E \approx \frac{\mathbb{I}_K}{K} \oplus_{E_S}\frac{\mathbb{I}_{K'_{E_S}}}{K'_{E_S}}
\otimes P_{E_S} \rho_S P_{E_S}
\end{aligned} \end{eqnarray}
We know \cite{uniqueinfo} that then any mixture of unitary transformations
can be performed on the system $R_2^ES^E$, provided $K$ is large with respect to $K'_E$,
and we shall choose $K$, and the size of the total system, in such a way,
that this is so, and at the same time $K'_E$ can be large too, which we will need further.
We shall below use notation $\eta'_{E-E_S}$ to denote the maximally mixed state acting on system $R_2^E$.
{\it Twirling} The following operation being a mixture of unitaries will prove useful.
For each fixed $E_S$ we apply a random unitary to $R_2^E$, and identity to the part $S^E$.
This operation does not change the final state of the system,
but greatly simplifies the form of the total state: namely for any initial
state on $R_2^ES^E$, the final state
is of the form
\begin{eqnarray} \begin{aligned}
\oplus _{E_S} \eta'_{E-E_S} \otimes \sigma_{E_S}
\label{eq:twirled}
\end{aligned} \end{eqnarray}
Finally let us note that we cannot perform any other operation, on the
state than a mixture of unitaries, because, for the total state of fixed energy block
\eqref{eq:eta_form} we can only apply some fixed unitary.
since in the process of tracing our over the reservoir,
we will sum over blocks, which effectively performs some mixture of unitaries.
However, the state with $\frac{\mathbb{I}}{K}$ tensored out does not actually differ much from
the state $\eqref{eq:eta_form}$, as we anyway will take the system $R_2^E$
to be large. Thus the output state coming from a mixture of unitaries performed on
the state with and without $\frac{\mathbb{I}}{K}$ being tensored out
have the same effect on the final form of the state of the system $S$.
We should now now recall, that the possibility of transforming one state into another by a convex combinations of unitary transformations
is simply given by the majorization conditions\cite{Uhlmann-ordering}.
The majorization condition reads as follows:
we have two sets of eigenvalues put in decreasing order $\{\lambda_i\}$ and $\{\lambda'_i\}$,
and we say that $\{\lambda_i\}$ majorizes $\{\lambda'_i\}$ when
\begin{eqnarray} \begin{aligned}
\sum_{i=1}^l \lambda_i \geq \sum_{i=1}^l\lambda_i'
\end{aligned} \end{eqnarray}
for all $l$. We say that $\rho$ majorizes $\sigma$ if the eigenvalues of $\rho$ majorize the
eigenvalues of $\sigma$.
In this way obtain the following theorem which will be the basis for our further results.
\begin{theorem}
\label{thm:major}
Consider two states $\rho_S$ and $\sigma_S$ diagonal in energy eigenbasis, on a system
with Hamiltonian $H_S$. The transition $(\rho_S,H_S) \otimes (\sigma_S,H_S)$ by means of thermal operations
is possible if and only if the state
\begin{eqnarray} \begin{aligned}
\oplus_{E_S}\eta_{E-E_S}\otimes P_{E_S} \rho_S P_{E_S}
\end{aligned} \end{eqnarray}
majorizes
\begin{eqnarray} \begin{aligned}
\oplus_{E_S}\eta_{E-E_S}\otimes \sigma_{E_S}
\end{aligned} \end{eqnarray}
for $E$ large enough. Moreover, if the above majorization relation holds
for two states $\rho_S$ and $\sigma_S$ not necessarily diagonal in energy eigenbasis,
then there exists $\sigma_S'$ such that for all $E_S$
$P_{E_S}\sigma_S P_{E_S}=P_{E_S}\sigma_S' P_{E_S}$, and the transition $(\rho_S,H_S) \to (\sigma_S',H_S)$
is possible.
\end{theorem}
Note that in the proof we have used the assumptions (i-iii) about the heat bath but not (iv).
The latter will be used when we will need to get rid of the heat bath in the majorization expressions.
Finally, it is intuitively obvious, that if we add to a heat bath a small system in a Gibbs state,
this is again a larger heat bath, i.e. it still satisfies our assumptions.
Indeed, consider a heat bath $R$ which satisfies assumptions (i-iv),
and another system $S'$, and consider the total Hamiltonian being a sum of Hamiltonians $H_R$ +$H_{S'}$.
The tensor product of two Gibbs states is a Gibbs state of a total system.
Since the original heat bath is large, and our system is small,
then the conditions \ref{item:peaked}, \ref{item:g_exp} and \ref{item:matching} are obviously satisfied.
Then, writing
\begin{eqnarray} \begin{aligned}
g^{RS'}(E) =\sum_{E_{S'}} g^R(E-E_{S'}) g^{S'}(E_{S'})
\end{aligned} \end{eqnarray}
and using the property \ref{item:g-property} of $g^R$ one gets
\begin{eqnarray} \begin{aligned}
g^{RS'}(E)\approx g^R(E) Z_{S'}
\label{eq:g_r_sprim}
\end{aligned} \end{eqnarray}
where $Z_{S'}$ is the partition function for $S'$.
This implies, in particular, that $g^{RS'}$ also satisfies the condition \ref{item:g-property}.
This proves the following intuitively obvious lemma:
\begin{lemma}
\label{lem:add_tau}
Transition between $\rho_S\otimes \tau_{S'}\to \sigma_S\otimes \tau_{S'}$
is possible if and only if transition $\rho_S\to \sigma_S$ is possible.
\end{lemma}
Thus adding a system in a Gibbs state makes sense only, if we consider transition
between systems with different Hamiltonian. Then we bring in a system in a Gibbs state,
only in order to have that Hamiltonian in future processes e.g. we might then transform the Gibbs state
into another state which needed to have that Hamiltonian.
The conditions given thus far for state transformations are all that is needed to draw the full amount
of work from a state, or to form a state from a heat bath. This is done in sections \ref{sec:pure} and \ref{sec:work}.
For the remainder of this section, we continue with more general state transformations.
\subsection{Thermo-majorization}
\label{sec:thermo_major}
We shall now provide an efficient method of finding, whether a transition $(\rho,H) \to (\sigma,H)$
is possible, for states which commute with Hamiltonian $H$.
The condition of transformations of the diagonal part of a density matrix given by theorem \ref{thm:major}
in terms of majorization involves not only the state, but also the heat bath, hence it is not always
directly useful. We shall now express the condition given
by majorization in terms of the states of system $S$ themselves which will result in an efficient
algorithm to decide whether a transition between two diagonal states is possible or not. Essentially, we need to write the eigenvalues
of the state and heat bath, in terms of eigenvalues of only the state.
We shall assume that our input state and output states are diagonal
in their energy bases, however, even if they are not, the condition we derive determines possible transformations of the diagonal part
of the density matrix, thus the condition becomes necessary, but ceases to be sufficient.
Let $p_{E_S,g}$ be eigenvalues of $\rho$ and $q_{E_S,g}$ be eigenvalues of $\sigma$.
Then, due to proposition \ref{thm:major} and the condition \eqref{item:g-property}, the state
$P_E \rho_R\otimes \rho_S P_E$ after normalisation is close to the state having the following eigenvalues:
\begin{eqnarray} \begin{aligned}
e^{\beta E_S}\frac{p(E_S,g)}{g_R(E)}
\label{eq:eig}
\end{aligned} \end{eqnarray}
with multiplicity $g_R(E) e^{-\beta E_S}$, where $E_S$ runs over all
energies of the system, and $g$ runs over degeneracies.
Similarly, $P_E \rho_R\otimes \sigma_S P_E$ has eigenvalues $e^{-\beta E_S}\frac{q(E_S,g)}{g_R(E)}$
with the same multiplicity.
The eigenvalues are very small, and they are collected in groups, where they are the same,
hence the majorization amounts to comparing integrals.
If one puts eigenvalues into decreasing order, one obtains a stair-case like function,
and majorization in the limit will be to compare the integrated functions (which are then
piece-wise linear functions).
To see how it works, we need to put the eigenvalues in nonincreasing order.
The ordering is determined by the ordering of the quantities $e^{\beta E_S}p_{E_S,g}$.
This determines the order of $p(E_S,g)$ (which in general will not be decreasing order anymore).
We shall denote such ordered probabilities as $p_i$,
and the associated energy of the eigenstate as $E_i$.
E.g. $p_1$ is equal to the $p(E_S,g)$ such that $e^{\beta E_S}p(E_S,g)$ is the largest.
Note that for fixed $E_S$ the order is the same as order of $P_{E_S,g}$,
while for different $E_S$ it is altered by the Gibbs factor. We do the same for $\sigma$,
which results in $q_i$.
The eigenvalues are thus ordered by taking into account Gibbs weights:
\begin{eqnarray} \begin{aligned}
\underbrace{\frac{p_1 e^{\beta E_1}}{{d_E}}}_{{\text{multiplicity}\atop \approx d_E e^{\beta E_1}}}
\geq
\underbrace{\frac{p_2 e^{\beta E_2}}{{d_E}}}_{{\text{multiplicity}\atop \approx d_E e^{\beta E_2}}}
\geq \ldots
\end{aligned} \end{eqnarray}
where $d_E$ is a shorthand for $g_R(E)$.
We shall now ascribe to vector $\{p_i\}$ a function mapping interval $[0,Z]$ into itself.
On the $y$ axis, we put subsequent sums $\sum_{i=1}^l p_i$, $l=1,\ldots ,d$ where
$d$ is the number of all probabilities,
and on the $x$ axis, we put sums $\sum_{i=1}^l e^{-\beta E_i}$, with the final point being at $x=Z$.
This gives $d+1$ pairs: $(0,0), (p_1, e^{-\beta E_1}), (p_1+p_2, e^{-\beta E_1}+e^{-\beta E_2}), \ldots,
(Z,1)$. We join the points, and it will gives us a graph of a function, $f_p(x)$.
It is easy to see, that in the limit of large $g_R(E)$, the eigenvalues of $\rho$ majorize eigenvalues of $\sigma$
if and only if $f_p(x)\geq f_q(x)$ for all $x\in[0,Z]$. The described scheme is presented on figure
\ref{fig:thermo-maj-idea}.
\begin{figure}
\label{fig:thermo-maj-idea}
\centering
\includegraphics[width=15cm]{fig-thermomaj-appendix.eps}
\caption{{\bf Thermo-majorization}. I. Standard majorization: (a) the histograms of probability distributions $\{p_i\}$ and $\{q_i\}$. (b) The distribution $\{p_i\}$ majorizes distribution $\{q_i\}$
if for all $l$ $\sum_{i=1}^l p_i\geq \sum_{i=1}^l q_l$. Graphically,this means that the entire
plot corresponding to $\{p_i\}$ is above the plot corresponding to $\{q_i\}$. II. Thermo-majorization.
(c) Here the histograms consist of groups of numerous columns of the same height. We set their base to $1/d_E$ where $d_E=g_R(E)$, which defines a stair-way looking function defined on interval $[1,Z]$,
where $Z$ is partition function. As a result, the plot analogous to that of (b), in the limit of
large $E$ (which implies $d_E\to \infty$) becomes integral of that function presented on panel (d).
The angles are given by $\tan(\alpha_i)=p_ie^{\beta E_i}$ hence they are decreasing. (e) checking thermo-majorization conditions amounts to comparing the plots which are piecewise linear functions.
The state $\rho_1$ thermo-majorizes each other state, while the thermal $\tau$ is thermo-majorized
by all other states. Thus we can transform $\rho_1$ into $\rho_2$, $\rho_3$ and $\tau$
and all states can be transformed into $\tau$. On the other hand, $\rho_2$ and $\rho_3$ are incomparable,
hence neither can be transformed into one another by Thermal Operations.}
\end{figure}
Note that the Gibbs state in this picture is represented by a trivial function $f_\beta(x)=Zx$
hence any state can be transformed into a Gibbs state. Note that one can generalise our new type of majorization, by replacing Gibbs state with an arbitrary state, obtaining an interesting mathematical generalisation of standard majorization. Likewise, although here the relevant conserved quantity is energy, one can generalise to operations which commute with any conserved quantity.
\section{Transitions involving pure excited states}
\label{sec:pure}
In preparation for deriving the expression for extracting work from a resource, or forming a state from the thermal state
by adding work, we will derive the condition for transitions involving a pure energy eigenstate. In particular, we will derive
the expression for extracting a pure excited state, and the expression for forming a state from a pure excited state. Then in
Section \ref{sec:work}, we will use the results in this section to derive our two free energies.
\subsection{Distillation: extracting a pure excited state}
\label{sec:dist}
In this section we derive the condition for when a given mixed state $\rho_S$ with Hamiltonian $H_S$
can be transformed into a pure excited state $\psi_W$ - an eigenstate of the
Hamiltonian $H_{S'}$ with eigenvalue $W$.
Let us first consider the case where we wish to extract $\psi_W$ with no probability of failure, from
a state diagonal in the energy basis. We will then extend our result to arbitrary states.
According to Lemma \ref{lem:add_tau} we need
to take an initial state $\rho_S\otimes \tau_{S'}$ and the final state
is an arbitrary state of the system $SS'$ of the form $\sigma_S\otimes|\psi^W\>\<\psi^W|_{S'}$.
Due to Theorem \ref{thm:major}, and Eq. \eqref{eq:g_r_sprim} a transition is possible when the state
\begin{eqnarray} \begin{aligned}
\bigoplus_{E_S}\eta_{E-E_S}^{RS'}\otimes P_{E_S} \rho_S P_{E_S}
\label{eq:in}
\end{aligned} \end{eqnarray}
majorizes
\begin{eqnarray} \begin{aligned}
\bigoplus_{E_S}\eta_{E-E_S-W}^R\otimes P_{E_S} \sigma P_{E_S}\otimes |\psi_W\>_{S'}\<\psi_W|
\label{eq:out}
\end{aligned} \end{eqnarray}
However, since $\sigma$ is arbitrary, and the target state of $S'$ is pure,
this is equivalent to the condition
\begin{eqnarray} \begin{aligned}
{\rm rank}_{in}\geq {\rm rank}_{out}.
\label{eq:rk}
\end{aligned} \end{eqnarray}
where ${\rm rank}_{in}$ and ${\rm rank}_{out}$ are ranks of the state \eqref{eq:in} and \eqref{eq:out},
respectively.
The rank of the initial state is equal to
\begin{eqnarray} \begin{aligned}
{\rm rank}_{in}=\sum_{E_S} g_{RS'}(E-E_S) {\rm rank}_{E_S} (\rho_S)
\end{aligned} \end{eqnarray}
where ${\rm rank}_{E_S} (\rho)$ is the rank of $P_{E_S} \rho_S P_{E_S}$,
and as in Eq. \ref{eq:g_r_sprim}
\begin{eqnarray} \begin{aligned}
g_{RS'}(E-E_S)=g_R(E-E_S)Z_{S'}.
\label{eq:g_r_sprim2}
\end{aligned} \end{eqnarray}
The maximal rank of the target state is given by
\begin{eqnarray} \begin{aligned}
{\rm rank}_{out}=\sum_{E_S} g_R(E-E_S-W) g_{S}(E_S)
\end{aligned} \end{eqnarray}
Now, using \eqref{eq:g_r_sprim2} and $g_R(E+\Delta E) \approx g_R(E) e^{\beta \Delta E}$ we obtain that
eq. \ref{eq:rk} implies
\begin{eqnarray} \begin{aligned}
\sum_{E_S}\frac{e^{-\beta(E_S)}}{Z}{\rm rank}_{E_S}(\rho_S) \leq \sum_{E_{S'}}\frac{e^{-\beta(E_{S'})}}{Z'}{\rm rank}_{E_{S'}}(\psi_W)
\end{aligned} \end{eqnarray}
which can be written as
\begin{eqnarray} \begin{aligned}
D_{\min}(\rho_S||\tau_S) \geq D_{\min}(\psi^W_{S'}||\tau_{S'})
\label{eq:dmincondition}
\end{aligned} \end{eqnarray}
with $D_{\min}(\rho_S||\tau_S):=-\ln\mathop{\mathsf{tr}}\nolimits{\Pi_{\rho}\tau}$. In general this quantity is the min-relative entropy\cite{datta2009min}.
We can now ask about the case when $\rho$ is not diagonal in the energy eigenbasis. In such a case, we simply replace $\rho$ with
$\omega=\sum_{Eg,g'}\proj{E,g}\rho_\epsilon\proj{E,g'}$ in Equation (\ref{eq:dmincondition}).
The reason, is that Theorem \ref{thm:major} states necessary and sufficient conditions
for transforming the diagonal entries of one density matrix into the diagonal entries of another.
In the case of an initial state with off-diagonal entries, it gives necessary conditions.
However the diagonal entries of a pure excited energy eigenstate determines uniquely that state itself, thus the condition must also
be sufficient. An alternative argument in terms of commuting of the dephasing operation and thermal operations is given in the Main Section.
Note that the operation which gets implemented to map one state to another is simply a mapping from eigenstates of the initial state
within each energy block $E$, to mappings of eigenstates of the final state within the same energy block. However, any such mapping will do, and
there are a huge number of them. Thus the experimenter does not need to know which unitary she is implementing, provided that it conserves energy.
She thus needs very little control over her systems -- she simply chooses any unitary which maps the macroscopic variables of one state (in this
case, total energies $(E_R,E_S)$, to macroscopic variables of the final state (in this case, a pure energy eigenstate
with no degeneracy on some system, and total energy on another $(E_R+E_S-W)$. The same is true of the formation process described in the next section.
\subsection{Formation of a resource state from a thermal bath and pure excited state}
\label{sec:form}
Just as one can draw work from a state which is out of equilibrium from the rest of the thermal bath,
it is also possible to perform the reverse process -- create a state from the thermal bath by adding work.
Here we provide conditions for transition from a pure excited state to a given target diagonal state. We will then
use it in Section \ref{sec:work} to derive the amount of work which is required to create a state.
We thus take the initial state to be of the form
\begin{eqnarray} \begin{aligned}
\rho^{in}=\psi^W_S\otimes \tau_{S'}
\end{aligned} \end{eqnarray}
and the output state
\begin{eqnarray} \begin{aligned}
\rho^{out} =\rho_{SS'}
\end{aligned} \end{eqnarray}
We shall now use Theorem \ref{thm:major}.
To this end we have to check the majorization condition between the following states:
\begin{eqnarray} \begin{aligned}
\eta_{E-W}^{RS'}\otimes |\psi^W\>_{S}\<\psi^W|
\label{eq:form_initial}
\end{aligned} \end{eqnarray}
and
\begin{eqnarray} \begin{aligned}
\oplus_{E_{S'}} \eta_{E-E_{S'}}^{RS}\otimes p_{E_{S'}}\rho^{E_{S'}}_{S'}
\end{aligned} \end{eqnarray}
where in \eqref{eq:form_initial} we have used Eq. \eqref{eq:g_r_sprim}.
However, the former state has only one eigenvalue $1/g_{RS'}(E-W)$ with multiplicity $g_{RS'}(E-W)$.
Therefore, the majorization condition is that all eigenvalues of the latter state
are no greater than this eigenvalue. I.e. we need that
\begin{eqnarray} \begin{aligned}
g_{RS'}(E-W)^{-1}\geq g_{RS}(E-E_{S'})^{-1} \lambda^{\max}_{E_{S'}}
\label{eq:major}
\end{aligned} \end{eqnarray}
holds for all $E_{S'}$, where $\lambda^{\max}_{E_{S'}}$ is the maximal eigenvalue of
$P_{E_{S'}} \rho_{S'} P_{E_{S'}}$ i.e. it is the maximal eigenvalue of $\rho_{S'}$
in the subspace of energy $E_{S'}$. Since the Hamiltonian for $RSS'$
is the sum of $H_R$, $H_S$ and $H_{S'}$, we obtain that
\begin{eqnarray} \begin{aligned}
&&g_{RS'}(E-W)=\sum_{E_{S'}} g_R(E-W-E_{S'}) g_S'(E_{S'}) \nonumber \\
&&g_{RS}(E-E_{S'})= \sum_{E_S} g_R(E-E_{S'}-E_S) g_S(E_S)
\end{aligned} \end{eqnarray}
Now we use the fact that $R$ is a heat bath, and we apply our assumption \eqref{eq:g-property}
which says that
\begin{eqnarray} \begin{aligned}
g_R(E-W-E_{S'})\approx g_R(E) e^{-\beta ( W +E_{S'})}
\end{aligned} \end{eqnarray}
and
\begin{eqnarray} \begin{aligned}
g_R(E-E_{S'}-E_S) \simeq g_R(E) e^{-\beta ( E_S +E_{S'})}
\end{aligned} \end{eqnarray}
we can thus rewrite the majorization condition \eqref{eq:major}
as follows
\begin{eqnarray} \begin{aligned}
\frac{1}{Z_{S'}}e^{-\beta E_{S'}} \geq \frac{1}{Z_S}e^{-\beta W} \lambda^{\max}_{E_{S'}}
\end{aligned} \end{eqnarray}
for all $E_{S'}$.
On the other hand, one can compute that
\begin{eqnarray} \begin{aligned}
&&D_{\max}(\psi^W_S||\tau_S) = Z_S e^{\beta W},
&&D_{\max}(\rho_{S'}||\tau_{S'}) = \max_{E_{S'}} Z_{S'} e^{\beta E_{S'}} \lambda^{\max}_{E_{S'}}
\end{aligned} \end{eqnarray}
where $D_{\max}(\rho||\tau):=\log\min \{\lambda:\rho\leq\lambda\tau \}$ is the max-relative entropy\cite{datta2009min}.
Thus, the transition $(\psi^W_S,H_S) \to (\rho_{S'},H_{S'})$ is possible if and only if
\begin{eqnarray} \begin{aligned}
D_{\max}(\psi^W_S||\tau_S) \geq D_{\max}(\rho_{S'}|\tau_{S'}).
\label{eq:dmaxcondition}
\end{aligned} \end{eqnarray}
\section{Extractable work, and work of formation}
\label{sec:work}
We now use the results of Secs. \ref{sec:form} and \ref{sec:dist} to discuss the amount of work
that can be drawn from a system in contact with a heat bath of temperature $T$, and the amount of work
that is needed to create one. In thermodynamics, both quantities are equal and are given by free energy.
In our case we obtain two free energies, $F_{min}$ governing extracting work,
and the other, $F_{max}$, governing creation of the system. In this section, we derive the expression for $F_{min}$ and $F_{max}$ in the case where we wish to extract the full amount of work available, or create a total state out of thermal states.
This corresponds to Equations \eqref{eq:ergotropy} and \eqref{eq:ergotropymax}.
The more general result of Equations \eqref{eq:gmin} and \eqref{eq:gmax} following from thermo-majorization is contained in the Main Section.
We propose to define the process of drawing or spending work as raising or lowering the energy level
of an eigenstate of a Hamiltonian $\hat{W}$ of a system. This system is used to store the energy provided
by drawing work. Thus we draw work $\Delta W$
if we transform a state $|E\>$ into $|E'\>$ such that $E'-E=W$ $H_W|E\>=E$ and $H|E\>=E$.
Expending work, would mean the reverse process. Since our results don't depend on the system used to store work,
we take the most elementary system than can be used, namely a two level system with energy gap $W$.
Thus consider a system $S$ in state $\rho_S$. We add a work system with Hamiltonian $\hat{W}$
in a state $|E\>$. Our initial state is thus $\rho_S\otimes \proj{E}$ and the final state
$\proj{E'}$. Using the results of section \ref{sec:dist} we obtain, that
$\rho_S\otimes \proj{E}$ can be transformed into $\proj{E'}$ if and only if
\begin{eqnarray} \begin{aligned}
D_{\min}(\rho_S\otimes \proj{E})\geq D_{\min}(\proj{E'}),
\label{eq:drawing-work}
\end{aligned} \end{eqnarray}
where we use the shorthand notation
$D_{\min}(\rho) \equiv D_{\min}(\rho|\tau)$. Since $D_{\min}$ is additive,
and for energy eigenstates $|E\>$ we have
\begin{eqnarray} \begin{aligned}
D_{\min}(|E\>)= \beta E -\ln Z_W
\end{aligned} \end{eqnarray}
where $Z$ is the partition function of the work system,
we can rewrite \eqref{eq:drawing-work} as
\begin{eqnarray} \begin{aligned}
k T D_{\min}(\rho_S)\geq W
\end{aligned} \end{eqnarray}
This allows us to define the free energy $F_{min}$ as follows:
\begin{eqnarray} \begin{aligned}
F_{min}= F_\beta + kT D_{min}
\end{aligned} \end{eqnarray}
where $F_\beta$ is the standard free energy of the equilibrium state (we have anyway that for thermal states $F_{min}=F_\beta$).
The work that can be drawn from a
non-equilibrium state is thus equal to the the free energy difference $\Delta F_{min}$:
\begin{eqnarray} \begin{aligned}
W_{dist}(\rho) = F_{min}(rho) - F_{min}(tau)
\end{aligned} \end{eqnarray}
We now wish to allow some probability of failure\cite{dahlsten2011inadequacy} --
namely, we might not produce $\psi^W_{S'}$ exactly, but rather a state $\psi^W_{\epsilon}$
$\epsilon$-close to $\psi^W$ i.e. such that
\begin{align}
||\psi^W_{\epsilon}-\psi^W||\leq \epsilon .
\label{eq:epsilonclose}
\end{align}
Since thermal operations are completely positive trace preserving maps,
then if we imagine the same operations were performed on some initial state $||\rho_{\epsilon}-\rho||\leq \epsilon$ then the final state
will also satisfy Equation (\ref{eq:epsilonclose}). We can thus replace $D_{min}(\rho||\tau)$ with
$D^\epsilon_{min}(\rho||\tau):=-\sup_{\rho_\epsilon} -\ln\mathop{\mathsf{tr}}\nolimits{\Pi_{\rho}\tau}$. This is known as the smooth-min entropy\cite{datta2009min}.
Analogously we define work which is needed to create a system, i.e.
we consider a transition $\proj{E}\to \proj{E'} \otimes \rho_S$, and in an analogous way obtain
that the minimal work $W=E'-E$ to ensure this transition is given by
\begin{eqnarray} \begin{aligned}
W_{form}(\rho) = F_{max}(\rho) - F_{max}(\tau)
\end{aligned} \end{eqnarray}
where $F_{max}$ is a max-free energy given by
\begin{eqnarray} \begin{aligned}
F_{max}=F_\beta + kT D_{max}.
\end{aligned} \end{eqnarray}
This comes from simply solving Equation \eqref{eq:dmaxcondition} for the value of $W$ required for the transition, to obtain
\begin{align}
W=
kT\inf_{\rho_{\epsilon}}
\log\min \{\lambda:\rho\leq\lambda\tau \}
\end{align}
As in the distillation process, we can consider $\epsilon$-close formation processes, since we will generally want to tolerate some small
error in the creation of a resource, particularly if it can save us needing a large amount of work.
We thus obtain the expression for $F_\epsilon^{max}$ in the Main Section.
\section{Characterisation of Thermal Operations}
We have provided an algorithm for deciding whether a state can be transformed into another state,
given by thermo-majorization. However the algorithm does not tell us what kind of
operations (completely positive maps) we can perform by means of Thermal Operations.
Below we shall show, that all possible processes are precisely those that preserve the Gibbs state.
This implies that if we have reversibility of state transformations (as is the case when we have many copies of a state\cite{thermoiid}), then the unique measure which determines whether a transformation is possible, is given by the relative entropy distance to the Gibbs state\cite{thermo-ent2002}. This quantity is the difference between free energy of a state of interest and that of Gibbs state \cite{donald1987free}. However, here, we do not have reversibility, thus there are at least two inequivalent functions which are non increasing under thermal operations ($F_{min}$ and $F_{max}$).
We start with a state $\tau_R\otimes\rho_S$ and write
\begin{eqnarray} \begin{aligned}
\tau_R\otimes\rho_S \approx \sum_{E\in {\cal E}} p_E \rho_{RS}^E
\end{aligned} \end{eqnarray}
with
\begin{eqnarray} \begin{aligned}
\rho_{RS}^E= \frac{1}{P_E}P_E\tau_R\otimes\rho_S P_E
\end{aligned} \end{eqnarray}
where ${\cal E}$ consists of very large energies in comparison with system energies,
and $p_E=\mathop{\mathsf{tr}}\nolimits(\tau_R\otimes\rho_S P_E)$.
We shall now fix one energy block, and show, that even when restricting just to
permutations of basis vectors within the block (being products of eigenstates of $\tau_R$ to eigenvalues $E_R$ and eigenstates of $\rho_S$ to eigenvalues $E_S$, such that $E_R+E_S=E$) we
can perform arbitrary operation on system $S$ which preserve the Gibbs state.
Then, we will argue that the operation on the system $S$ can be made the same
for each energy block (for $E\in {\cal E}$).
To prove the first claim, for simplicity,
let us assume that the Hamiltonian $H_S$ is nondegenerate (extension to the degenerate case is immediate).
As follows from Theorem \ref{thm:major}, in such a fixed subspace, the eigenvalues of
our state form groups labelled by energy $E_S$. Within each group, we have
$g_R(E) e^{-\beta E_S}$ eigenvalues all equal to $\frac{p(E_S)}{g_R(E)e^{-\beta E_S}}$.
Permutations of basis vectors result in transferring some subsets of a given group to other groups.
Let us then use indices $i$ in place of $E_S$, so that $p_{E_S}\to p_i$ and $g_R(E)e^{-\beta E_S}\to d_i$.
We shall denote by $k_{i\to j}$ the "transition current" i.e. the number of eigenstates that have been moved from the $i$-th group to the $j-th$ group.
Clearly $k_{i\to j}$ satisfy
\begin{eqnarray} \begin{aligned}
\sum_j k_{i\to j} =d_i\\
\label{eq:k1}
\quad \sum_i k_{i\to j}=d_j\\
\label{eq:k2}
\end{aligned} \end{eqnarray}
The transition "currents" are illustrated on Fig. \ref{fig:processes}.
\begin{figure}
\centering
\includegraphics[width=10cm]{processes.eps}
\caption{Transitions for two level system with energy levels $E=0$ and $E=1$. a) There are four possible transitions b) the number of states of bath corresponding to each level is proportional
to the Boltzmann factor.}
\label{fig:processes}
\end{figure}
After an operation given by some fixed set of $k_{i\to j}$ satisfying the above transitions, we obtain
a new state, whose probabilities $q_i$ are given by
\begin{eqnarray} \begin{aligned}
q_j=\sum_i k_{i\to j} \frac{p_i}{d_i}
\end{aligned} \end{eqnarray}
Thus, we can define transition probabilities $p_{i\to j}$ as
\begin{eqnarray} \begin{aligned}
p_{i\to j}=\frac{k_{i\to j}}{d_i}
\end{aligned} \end{eqnarray}
Then the condition \eqref{eq:k1} means that the $p_{i\to j}$ ensure normalisation, so
that the only constraint on possible process is \eqref{eq:k2}. However, since
$\frac{d_i}{d_j}=\frac{e^{\beta E_j}}{e^{\beta E_j}}$, the latter condition
means simply that the Gibbs state is preserved. This ends the proof,
that for fixed $E$ we can perform all Gibbs preserving operations.
Finally, given arbitrary Gibbs preserving transformation on $S$
we perform for every total energy block permutation
that results in this transformation. In this way the needed transformation is
performed on the initial state of system $S$. Of course, Thermal Operations obviously
do preserve Gibbs state, hence we obtain, that Thermal Operations are arbitrary operations
that preserve Gibbs state.
Let us discuss this result in the context of the detailed balance condition. The latter
is the property that $\frac{p_{i\to j}}{p_{j\to i}}=e^{-\beta (E_j-E_i)}$.
As we will see, Thermal Operations need not satisfy detailed balance; they should merely
preserve the Gibbs state as a whole. To provide an example, let us distinguish a class of Gibbs-preserving processes called quasi-cycles:
we put the energy levels on a circle, and from one level, one can go only to the next neighbouring level, as in Figure \ref{fig:quasi-cycles}
\begin{figure}
\centering
\includegraphics[width=12cm]{quasi-cycles.eps}
\caption{Quasi-cycles. a) In quasi cycle for each level there is only one transition to a different level b) quasi-cycle for
three level system. The points represent levels with energies $0,1$ and $2$; probability $p_{2\to2}$
of staying in level $2$ vanishes. c) Currents for three-level quasi-cycle: the shaded microstates
are subjected to a cycle, while the ones not shaded are left untouched.}
\label{fig:quasi-cycles}
\end{figure}
The simplest description of a quasi-cycle is in terms of
quantities $k_{i\to j}$. Namely, we choose an order of levels, put them on a circle, fix
a direction, and the process is to take all states from the group
of states with the largest energy $E_S$, and shift them to the states with the energy level in the chosen direction.
I.e. the process is determined by $p_{i\to i+1}=e^{-\beta (E_{\max}-E_i)}$,
where $E_{\max}$ is the maximal energy, and $E_i$ is the energy of $i-th$ level.
For two level systems, the class of Gibbs preserving operations is the same as the class
of operations satisfying the detailed balance condition,
and all possible processes are parametrised by a single number $r\in[0,1]$, which is the probability of
mixing two basic processes: the identity operation, and the two-level quasi-cycle.
For three level systems, there are processes that preserve the Gibbs state,
but do not satisfy detailed balance, an example being the three-level quasicycle. It turns out that the class of Gibbs preserving
maps is strictly more powerful that the class of detailed-balance maps. An example is the transition
between $(0,\frac12,frac12)$ and $(\frac12e^{-2\beta}, \frac12(1- e^-\beta),\frac12(e^{-\beta}-e^{-2\beta}+1)$
with the energy levels given by $(0,1,2)$. It turns out that the only Gibbs preserving operation that can transform the first state into the second one is the quasi-cycle $0\to1\to2$.
This means that such a transition is impossible by means of weak coupling with the heat-bath
\cite{Alicki-Lendi}, as at weak coupling the detailed balance condition is satisfied.
\end{appendix}
\end{document}
|
\section*{Introduction}
Throughout this paper, algebraic cycles and Chow groups are with
rational coefficients. Let $X$ be a smooth projective complex
variety. The Hodge conjecture predicts that every Hodge class in
$H_{2i}(X) := H_{2i}(X(\C),\Q)$ is the class of an algebraic cycle. In
particular, given any two smooth projective complex varieties $X$ and
$Y$, the Hodge conjecture predicts that any morphism $f$ of Hodge
structures between $H_*(X)$ and $H_*(Y)$ comes from geometry. By this
we mean that $f$ is induced by a correspondence between $X$ and $Y$,
that is, by an algebraic cycle on $X \times Y$. Whether the Hodge
conjecture happens to be true or not, Grothendieck pointed out that
certain morphisms of Hodge structures play a more important role in
the theory of algebraic cycles. If $X$ has pure dimension $d$, he
suggested that for all integer $i$ the K\"unneth component in
$H_{2d}(X \times X)$ inducing the projector on $H_i(X)$ should be
induced by a correspondence. He also suggested that, for all $i\leq
d$, the inverse to the Lefschetz isomorphism $H_{2d-i}(X) \r H_i(X)$
given by intersecting $d-i$ times with the class of a smooth
hyperplane section should be induced by an algebraic cycle. The first
conjecture is usually referred to as the K\"unneth standard conjecture
and the second one to the Lefschetz standard conjecture. Classically
\cite[4.1]{Kleiman}, it is known that the Lefschetz standard
conjecture for $X$ implies the K\"unneth standard conjecture for $X$.
If the Hodge conjecture gives a simple description of the image of the
cycle class map $\CH_i(X) \r H_{2i}(X)$ in terms of the Hodge structure
of $H_{2i}(X)$, it is a much more difficult problem to unravel the
nature of its kernel. Beilinson and Bloch, inspired by Grothendieck's
philosophy of motives, first proposed a description of such a kernel
in terms of a descending filtration on Chow groups that would behave
functorially with respect to the action of correspondences and would
be such that its graded parts would depend solely on the homological
motive of $X$.
More generally, the conjectures of Bloch and Beilinson can be
formulated for smooth projective varieties defined over any field $k$
if one uses $\ell$-adic cohomology in place of Betti cohomology. In
that setting, for $X$ smooth projective over $k$, the K\"unneth and
Lefschetz standard conjectures stipulate the existence of algebraic
cycles in $X \times X$ inducing the right action on cohomology. This
is consistent with the previous formulations. Indeed, let $X$ be a
smooth projective variety defined over a subfield $k \subseteq \C$ and
assume that there is a cycle $\Gamma \in \CH^{i}(X_\C \times_\C X_\C)$
inducing the inverse to the Lefschetz isomorphism $H_{2d-i}(X) \r
H_i(X)$. Here, $d$ is the dimension of $X$ and $i$ is a non-negative
integer $\leq d$. It is a fact that, for a smooth projective variety
$Y$ defined over $k$, the base change map $\CH_*(Y_K) \r \CH_*(Y_L)$ is
an isomorphism modulo homological equivalence for all extensions of
algebraically closed fields $L/K$ over $k$. Therefore, up to replacing
$\Gamma$ by a cycle homologically equivalent to it, we can assume that
$\Gamma$ is defined over the algebraic closure of $k$ inside $\C$.
But then, $\Gamma$ is defined over a finite extension of $k$ which can be
chosen to be Galois, say of degree $n$, and with Galois group $G$. It
is then straightforward to check that the cycle $\gamma := 1/n \cdot
\sum_{\sigma \in G} \sigma_*\Gamma$ is defined over $k$ and induces
the inverse to the Lefschetz isomorphism $H_{2d-i}(X) \r H_i(X)$. The
same arguments apply to the K\"unneth standard conjecture.
Twenty years ago, Murre proposed that not only should the K\"unneth
projectors in cohomology be induced by correspondences, but also that
they should be induced by correspondences that are idempotents modulo
rational equivalence. Given $X$ a smooth projective variety of
dimension $d$ over a field $k$, Murre \cite{Murre1} conjectured the
following. \medskip
(A) $X$ has a Chow--K\"unneth decomposition $\{\pi_0, \ldots,
\pi_{2d}\}$ : There exist mutually orthogonal idempotents $\pi_0,
\ldots, \pi_{2d} \in \CH_d(X \times X)$ adding to the identity such
that $(\pi_i)_*H_*(X)=H_i(X)$ for all $i$.
(B) $\pi_0, \ldots, \pi_{2l-1},\pi_{d+l+1}, \ldots, \pi_{2d}$ act
trivially on $\CH_l(X)$ for all $l$.
(C) $F^i\CH_l(X) := \ker(\pi_{2l}) \cap \ldots \cap \ker(\pi_{2l+i-1})$
does not depend on the choice of the $\pi_j$'s. Here the $\pi_j$'s are
acting on $\CH_l(X)$.
(D) $F^1\CH_l(X) = \CH_l(X)_\hom$, where the subscript `hom' refers to
homologically trivial cycles. \medskip
A variety $X$ which has a Chow--K\"unneth decomposition that satisfies
Murre's conjectures (B) and (D) is said to have a \emph{Murre
decomposition}. If moreover the Chow--K\"unneth decomposition of
conjecture (A) can be chosen so that $\pi_i = {}^t\pi_{2d-i} \in
\CH_d(X \times X)$, then $X$ is said to have a \emph{self-dual Murre
decomposition}. Here, we understand Murre's conjecture (C) as saying
that any two filtrations induced by two distinct Chow--K\"unneth
decompositions for $X$ coincide, not that they are merely isomorphic.
The relevance of Murre's conjectures was demonstrated by Jannsen
\cite{Jannsen} who showed that these hold for all smooth projective
varieties if and only if Bloch's and Beilinson's conjectures hold for
all smooth projective varieties. Murre's formulation of a conjectural
descending filtration on Chow groups has the advantage over
Beilinson's and Bloch's that it does not involve any functoriality
properties and that it can therefore be proven on a case-by-case
basis. Since Murre's paper \cite{Murre1} appeared, many authors have
tried to prove those conjectures for certain classes of varieties. In
this paper, we extend the list of cases for which these can be proven.
\medskip
Let $X$ be a smooth projective variety defined over a field $k$ and
let $\Omega$ be a universal domain over $k$, i.e. $\Omega$ is an
algebraically closed field of infinite transcendence degree over $k$.
A smooth projective variety $X$ will be said to have trivial Chow
group of zero-cycles if $\CH_0(X_\Omega) = \Q$. Our main result is the
following
\begin{theorem2} \label{mainth} Let $f:X \r S$ be a dominant morphism
between smooth projective varieties defined over a field $k$ such
that the general fibre of $f_\Omega$ has trivial Chow group of
zero-cycles. Suppose that $S$ has dimension $\leq 2$ and that $X$
has dimension $\leq 4$. Then $X$ has a self-dual Murre
decomposition. Moreover, the motivic Lefschetz conjecture, as stated
in \S\ref{motLef}, holds for $X$ and hence so does the Lefschetz
standard conjecture.
\end{theorem2}
Let's stress that the theorem gives a self-dual Murre decomposition
of $X$ which is defined over a field of definition of $f$.
Together
with standard results on rationally connected varieties
\cite[IV.3.11]{Kollar} (see also the proof of Corollary
\ref{ratconncoro}), we deduce
\begin{theorem2} \label{ratconn2}
Let $f:X \r S$ be an equidimensional dominant morphism between
smooth projective varieties defined over a field $k$ whose general
fibre is separably rationally connected. Suppose that $S$ has
dimension $\leq 2$ and that $X$ has dimension $\leq 4$. Then $X$
has a self-dual Murre decomposition which satisfies the motivic
Lefschetz conjecture.
\end{theorem2}
Theorem \ref{mainth} contrasts with the approach of
Gordon--Hanamura--Murre \cite{GHM} where Chow--K\"unneth decompositions
are constructed for varieties $X$ that come with a fibration $f : X
\r S$ which is ``nice'' enough: it is assumed, among other things, in
\emph{loc. cit.} that $f$ should be smooth away from a finite number
of points on $S$ and that $f$ should have a relative Chow--K\"unneth
decomposition. Here, we do not even require $f$ to be flat.
Theorem \ref{mainth} was already proved in the case $\dim S \leq 1$ :
a self-dual Chow--K\"unneth decomposition for which the motivic
Lefschetz conjecture holds was constructed in \cite[4.6]{Vial3} and
Murre's conjectures were checked to hold in \cite[4.21]{Vial2}. The
results in \cite{Vial2, Vial3} just mentioned are more generally
valid for fourfolds with Chow group of zero-cycles supported on a
curve. By Theorem \ref{isogeny}, it is the case that, if $X$ is as in
Theorem \ref{mainth} with $\dim S \leq 1$, then $\CH_0(X_\Omega)$ is
supported on a curve. From now on, we will therefore focus on the
case $\dim S =2$.
Here, Murre's conjecture (C) is proved only on the grounds that the
idempotents of a Chow--K\"unneth decomposition for $X$ are supported
in a specific dimension, cf. section \ref{Murreconj}. It is however
fully proved under some extra assumption on $X$ in Theorem
\ref{Murre-theorem} and Theorem \ref{C-Murre2}. Let's mention that
del Angel and M\"uller-Stach \cite{dAMS} proved the existence of a
Murre decomposition for threefolds fibred by conics over a surface
(see also the recent paper \cite{MSS} where Chow--K\"unneth
decompositions are constructed for some threefolds including conic
fibrations). In addition to treating the four-dimensional case, our
theorem makes more precise the result of \cite{dAMS} by showing that
the Murre decomposition can be chosen to be self-dual and by showing
the motivic Lefschetz conjecture for $X$. Also our approach is
different from \cite{dAMS}. Del Angel and M\"uller-Stach assume that
all the fibres of $f$ are rationally connected, this allows them to
compute the cohomology of $X$ via the Leray spectral sequence. They
then construct idempotents modulo rational equivalence and check that
they act as the K\"unneth projectors on cohomology. Here, we do not
make any assumptions on the bad fibres of $f$ and we first compute
the Chow group of zero-cycles of $X$ to only then deduce that the
idempotents we construct act as the K\"unneth projectors on
cohomology. \medskip
A word about the proof of the theorem and about the organisation of
the paper. In section \ref{geom}, we make the simple observation that,
for $f$ as in Theorem \ref{mainth}, $f_* : \CH_0(X_\Omega) \r
\CH_0(S_\Omega)$ is bijective with inverse induced by an algebraic
correspondence which is defined over a field of definition of $f$.
Together with Theorem \ref{effective}, the proof of the existence of
a Murre decomposition for $X$ essentially reduces to the case of
motives of surfaces. The validity of Murre's conjectures (A), (B) and
(D) for surfaces goes back to Murre himself \cite{Murre}. However,
by the very nature of Theorem \ref{effective}, we need Murre's
conjectures not only for surfaces but for motives of surfaces (i.e.
we need to deal with idempotents). This is the object of section
\ref{AP}. The construction of idempotents inducing the right
K\"unneth projectors for $X$ in homology is carried out in section
\ref{CK}. Constructing such idempotents is easy from the case of
surfaces. However, these are not necessarily mutually orthogonal. The
non-commutative Gram--Schmidt process which already appears in
\cite{Vial3} and which is run on this set of idempotents is described
in Lemma \ref{linalg}. This way, we obtain a \emph{self-dual}
Chow--K\"unneth decomposition for $X$. The motivic Lefschetz
conjecture is formulated in section \ref{motLef}, its relevance is
discussed and it is proved for $X$ there. Murre's conjectures (B)
and (D) are then proved for $X$ in section \ref{Murreconj} by using
results of \cite{Vial2} which are recalled in Proposition
\ref{MurreB} and \ref{MurreC}. If $C$ is a smooth projective curve,
then the results of \emph{loc. cit.} actually make it possible to prove
Murre's conjectures (B) and (D) for $X \times C$. This is the object
of section \ref{Murreconj2}.
Murre's observation that the K\"unneth projectors should lift to
idempotents modulo rational equivalence is crucial in the sense that
a combination of Beilinson's and Bloch's conjectures with
Grothendieck's standard conjectures imply that any projector in
homology should be liftable to an idempotent modulo rational
equivalence. Shun-Ichi Kimura \cite{Kimura} introduced a notion of
finite-dimensionality for Chow motives which implies such a lifting
property for projectors. Kimura's notion of finite-dimensionality has
become widely popular for this reason and more importantly because of
its relationship to Murre's conjectures. The simple observation of
Theorem \ref{isogeny} is used in section \ref{fdprob} to give new
examples of threefolds of general type which are Kimura
finite-dimensional, namely threefolds fibred by Godeaux surfaces.
There, using the result of section \ref{CK}, we also show in Theorem
\ref{conic-Kimura} that, if $X$ is a conic fibration over a surface
which is Kimura finite-dimensional, then $X$ is Kimura
finite-dimensional. In section \ref{general}, we produce examples of
fourfolds of general type with Chow group of zero-cycles not
supported on a curve but which admit a self-dual Murre decomposition.
Such examples will be given by fourfolds fibred by surfaces
birational to Godeaux surfaces. Theorem \ref{isogeny} is slightly
generalised in Theorem \ref{repsurface} ; this is used in section
\ref{unram} to prove finiteness of unramified cohomology in some new
cases. \medskip
Finally, although we don't state it here, the methods of this paper
actually show that Murre's conjectures hold for smooth projective
fourfolds $X$ for which there exist a smooth projective surface $S$
and correspondences $\alpha \in \CH_2(S \times X)$ and $\beta \in
\CH^2(X \times S)$ such that $\beta \circ \alpha = \Delta_S \in \CH_2(S
\times S)$ and such that $\alpha \circ \beta$ acts as the identity on
$\CH_0(X_\Omega)$. Consequently, if $f : X \r S $ is a dominant
morphism to a surface $S$ such that the general fibre of $f_\Omega$
has trivial Chow group of zero-cycles, then Murre's conjectures hold
for any smooth projective variety $X'$ which is birational to $X$.
\paragraph{Notations.} We refer to \cite{Scholl} for the notion of
Chow motive and to \cite{KMP} for the covariant notations we use
here. Briefly, a Chow motive $M$ is a triple $(X,p,n)$ where $X$ is
a smooth projective variety over $k$ of pure dimension $d$, $p \in
\CH_d(X\times X)$ is an idempotent ($p\circ p = p$) and $n$ is an
integer. The motive $M$ is said to be effective if $n \geq 0$. The
dual of $M$ is the motive $M^\vee := (X,{}^tp,-d-n)$, where ${}^tp$
denotes the transpose of $p$. A morphism between two motives
$(X,p,n)$ and $(Y,q,m)$ is a correspondence in $q \circ \CH_{d+n-m} (X
\times Y) \circ p$. We write $\h(X)$ for the motive of $X$, i.e. for
the motive $(X,\Delta_X,0)$ where $\Delta_X$ is the class of the
diagonal inside $\CH_d(X \times X)$. We have $\CH_i(X,p,n) =
p_*\CH_{i-n}(X)$ and $H_i(X,p,n) = p_*H_{i-2n}(X)$, where we write
$H_i(X) := H^{2d-i}(X(\C),\Q)$ for Betti homology when $k \subseteq
\C$. The results of this paper are valid more generally for any field
$k$ if one considers $\ell$-adic homology $H_i(X,\Q_\ell) :=
H^{2d-i}(X_{\bar{k}},\Q_\ell)$ with $\ell \neq \mathrm{char} \ k$ in
place of Betti homology. In that case the action of a correspondence
$\Gamma \in \CH_i(X \times Y)$ on $H_i(X,\Q_\ell)$ is given by the
action of $\Gamma \otimes_\Q 1 \in \CH_i(X \times Y) \otimes_\Q
\Q_\ell$.
\paragraph{Acknowledgements.} I am grateful to Sergey Gorchinskiy for
his interest and for providing me with a simpler construction of the
Albanese projector. Thanks to the referee for his detailed remarks
and numerous suggestions. Thanks to Mingmin Shen for stimulating
discussions and Burt Totaro for useful comments. This work is
supported by a Nevile Research Fellowship at Magdalene College,
Cambridge and an EPSRC Postdoctoral Fellowship under grant
EP/H028870/1. I would like to thank both institutions for their
support.
\section{A geometric result on zero-cycles} \label{geom}
Let $X$ and $S$ be smooth projective varieties over $k$ and let $f : X
\r S$ be a dominant morphism. Then any general linear section $\sigma
: H \rightarrow X$ of dimension $\dim S$ is smooth over $k$ and is
such that the morphism $\pi :=f|_H : H \r S$ is dominant. In
particular, for any general $H$, $\pi$ is generically finite and its
degree is written $n$.
\begin{proposition} \label{surjective} Let $f: X \rightarrow S$ be a
dominant morphism. Then, for any general $H$ as above, $\Gamma_f
\circ \Gamma_\sigma \circ {}^t\Gamma_\pi = n\cdot \Delta_S \in
\CH_{\dim S}(S \times S)$.
\end{proposition}
\begin{proof} This follows from the projection formula applied to $\pi
= f \circ \sigma$ and Manin's identity principle. See \cite[Example
1 p. 450]{Manin}.
\end{proof}
\begin{definition}
Let $f: X \rightarrow S$ be a dominant morphism between smooth
projective varieties defined over a field $k$. A general point of
$S$ is a closed point sitting outside a given proper closed subset
of $S$. By \emph{general fibre} of $f$, we mean the fibre of $f$
over a general point of $S$.
\end{definition}
\begin{theorem} \label{isogeny} Let $f: X \rightarrow S$ be a dominant
morphism between smooth projective varieties defined over a field
$k$. Assume that a general fibre $Y$ of $f$ satisfies $\CH_0(Y) =
\Q$.
Then the induced map $f_* : \CH_0(X) \r \CH_0(S)$ is bijective and
its inverse is induced by a correspondence $\Gamma \in \CH^{\dim X}(S
\times X)$. Moreover $\Gamma$ can be chosen to be defined over a
field of definition of $f$.
\end{theorem}
\begin{proof} Let's show that the correspondence $\Gamma$ can be
chosen to be $\frac{1}{n} \Gamma_\sigma \circ {}^t\Gamma_\pi$.
According to Proposition \ref{surjective}, it suffices to prove that
the correspondence $ \Gamma_\sigma \circ {}^t\Gamma_\pi$ induces a
surjective map $(\Gamma_\sigma \circ {}^t\Gamma_\pi)_* : \CH_0(S) \r
\CH_0(X)$.
Let's fix an open subset $U$ of $S$ such that $\pi : H_U \r U$ is
finite and such that the fibres of $f_U$ satisfy $\CH_0(X_u)=\Q$ for
all closed points $u$ in $U$.
Let $p$ be a closed point of $X$ and let $[p]$ denote the class of
$p$ in $\CH_0(X)$. By Chow's moving lemma, the zero-cycle $[p] \in
Z_0(X)$ is rationally equivalent to a zero-cycle $\alpha = \sum a_i
\cdot [p_i]$ supported on $X_U$. This means that each $p_i$ is a
closed point of $X$ that belongs to the open subset $X_U$ of $X$.
Let $u_i := f(p_i)$. Since $\CH_0(X_{u_i}) = \Q$ and $\deg
(\sigma_*\pi^*[u_i]) \neq 0$, we see that $p_i$ is rationally
equivalent to $\sigma_*\pi^*[u_i]$ taken with an appropriate
rational coefficient. Thus, $\sigma_*\pi^*$ is surjective on
zero-cycles.
\end{proof}
For example, we get as a corollary the following which is used in
\cite[Cor. 4.23]{Vial2} and \cite[Cor. 4.7]{Vial3} in the cases when
$S$ is a curve.
\begin{corollary} \label{ratconncoro} Let $f : X \rightarrow S$ be an
equidimensional dominant morphism between smooth projective
varieties defined over a field $k$. Assume that the general fibre of
$f$ is separably rationally connected (e.g. $X$ could be a Fano
fibration). Then $f_*:\CH_0(X_\Omega) \r \CH_0(S_\Omega)$ is
bijective and there is a correspondence $\Gamma \in \CH^{\dim X}(S
\times X)$ such that $\Gamma_* : \CH_0(S_\Omega) \r \CH_0(X_\Omega)$
is the inverse of $f_*$. \qed
\end{corollary}
\begin{proof}
According to Theorem \ref{isogeny}, it suffices to prove that a
general fibre of $f_\Omega$ has trivial Chow group of zero-cycles.
By considering a smooth closed fibre of $f$ which is separably
rationally connected, we see after pulling back to $\Omega$ that
$f_\Omega$ has a smooth closed fibre which is separably rationally
connected. It follows from \cite[IV.3.11]{Kollar} that a general
fibre of $f_\Omega$ is separably rationally connected and hence that
a general fibre of $f_\Omega$ has trivial Chow group of zero-cycles.
\end{proof}
Theorem \ref{repsurface} below, which generalises Theorem
\ref{isogeny} is irrelevant to the proof of Theorem \ref{mainth}.
However, we include it here because of the interesting consequences it
has for unramified cohomology, see section \ref{unram}.
\begin{definition}
A smooth projective variety $X$ over $k$ is said to have
\emph{representable} Chow group of algebraically trivial $i$-cycles
if there exists a curve $C$ over $\Omega$ and a correspondence
$\gamma \in \CH_{i+1}(C \times X_\Omega)$ such that
$\gamma_*\CH_0(C)_{alg} = \CH_i(X_\Omega)_{alg}$.
\end{definition}
\begin{lemma} \label{rep} Let $X$ be a smooth projective variety over
$k$. Then the following statements are equivalent.
$1.$ $\CH_0(X)_{\alg}$ is representable.
$2.$ The Albanese map $\mathrm{alb}_{X_\Omega} :
\CH_0(X_\Omega)_{alg} \r \mathrm{Alb}_{X_\Omega}(\Omega)$ is an
isomorphism (this map is always surjective).
$3.$ If $\iota :C \r X$ is any smooth linear
section of $X$ of dimension $1$, the induced map $\iota_* :
\CH_0(C_\Omega) \r \CH_0(X_\Omega)$ is surjective.
\end{lemma}
\begin{proof} This was proved by Jannsen \cite[1.6]{Jannsen}.
\end{proof}
\begin{theorem} \label{repsurface} Let $f : X \rightarrow S$ be a
generically smooth and dominant morphism between smooth projective
varieties defined over a field $k$. Assume that the general fibre
$Y$ of $f_\Omega$ is such that $\CH_0(Y)_\alg$ is representable.
Then $\CH_0(X)$ is supported in dimension $\dim S +1$. This means
that there exists a smooth projective variety $T$ over $k$ of
dimension $\dim S +1$ and a correspondence $\Gamma \in \CH^{\dim X}(T
\times X)$ such that $(\Gamma_\Omega)_* : \CH_0(T_\Omega) \r
\CH_0(X_\Omega)$ is surjective.
\end{theorem}
\begin{proof} Let $\iota : H \r X$ be a smooth linear section of $X$
of dimension $\dim S + 1$ such that $f$ restricted to $H$ is
dominant and generically smooth. Let $U$ be an open subset of
$S_\Omega$ such that $f_U : X_U \r U$ is smooth, $f_U|_{H_U} : H_U
\r U$ is smooth and such that the fibres of $f_U$ have representable
Chow group of zero-cycles.
Let's prove that $(\iota_\Omega)_* : \CH_0(H_\Omega) \r
\CH_0(X_\Omega)$ is surjective. Let $p$ be a closed point of
$X_\Omega$. By Chow's moving lemma, the zero-cycle $[p]$ is
rationally equivalent to a cycle $\alpha = \sum a_i \cdot [p_i]$
supported on $X_U$. Let $s_i := f(p_i) \in U$. Then, by choice of
$U$, each cycle $[p_i]$ is rationally equivalent on $X_{s_i}$ to a
cycle $\beta_i$ supported on $H_{s_i}$. Now clearly $\sum a_i \cdot
\beta_i$ is in the image of $(\iota_\Omega)_* : \CH_0(H_\Omega) \r
\CH_0(X_\Omega)$ and hence so is $[p]$.
\end{proof}
\begin{remark}
The descent properties of Theorems \ref{isogeny} \&
\ref{repsurface}, i.e. the fact that the correspondence $\Gamma$ in
those theorems can be chosen to be defined over a field of
definition of $f$, are essential to proving that $X$ has a
Chow--K\"unneth decomposition defined over the field of definition of
$f$ (Theorem \ref{theoremCK}) and to proving Proposition
\ref{bielliptic-unram}.
\end{remark}
\begin{remark} Under the assumptions of the above theorem, it is not
true that if $\CH_0(S)$ is supported in dimension, say $n$, then
$\CH_0(X)$ is supported in dimension $n+1$. Consider for example a
Lefschetz fibration $S \r \P^1$ associated to a smooth projective
surface $S$ with non-representable Chow group of zero-cycles.
\end{remark}
\begin{remark} \label{BB} Let $Y$ be a smooth projective variety over
$k$ and let $H \r Y$ be a smooth linear section of dimension $n$. A
consequence of the conjectures of Bloch and Beilinson is that, if
$\CH_0(Y_\Omega)$ is supported in dimension $n$, then $\CH_0(H_\Omega)
\r \CH_0(Y_\Omega)$ is surjective. Therefore, if one believes in the
conjectures of Bloch and Beilinson, then the above theorem can be
extended to the following. Let $f : X \r S$ be a generically smooth
and dominant morphism between smooth projective varieties defined
over a field $k$. Let $n$ be a positive integer and assume that the
general fibre $Y$ of $f$ is such that $\CH_0(Y)$ is supported in
dimension $n$. Then $\CH_0(X)$ is supported in dimension $\dim S +
n$.
\end{remark}
\section{Effective motives with trivial Chow group of zero-cycles}
The following theorem was mentioned to me by Bruno Kahn. Its proof
uses, among other things, the technique of Bloch and Srinivas
\cite{BS} together with Theorem 2.4.1 of Kahn--Sujatha
\cite{KahnSujatha} where it is shown that a correspondence $\Gamma \in
\CH_d(X \times Y) = \Hom(\h(X),\h(Y))$ which vanishes in $\CH_d(U \times
Y)$ for some open subset $U \subset X$ factors through some motive
$\h(Z)(1)$ with $\dim Z = d -1$.
\begin{theorem}\label{effective}
Let $M=(X,p)$ be an effective Chow motive such that $\CH_0(M_\Omega)
= 0$. Then there exist a smooth projective variety $Y$ of
dimension $\dim X -1$ and an idempotent $q \in \CH_{\dim Y}(Y \times
Y)$ such that $(X,p,0) \simeq (Y,q,1)$.
\end{theorem}
\begin{proof} Without loss of generality, we can assume that $X$ is
connected. Since we are working with rational coefficients, the
assumption $\CH_0(M_\Omega)=0$ implies $\CH_0(M_{k(X)})=0$ which means
$p_*\CH_0(X_{k(X)}) = 0$. In particular, if $\eta_X$ denotes the
generic point of $X$, then we have $p_*\eta_X=0$. But $p_*\eta_X$
is the restriction of $p \in \CH_d(X \times X)$ to $\varinjlim \CH_d(U
\times X) = \CH_0(X_{k(X)})$, where the limit is taken over all open
subsets $U$ of $X$. Therefore, by the localization exact sequence
for Chow groups, there exist a proper closed subset $D \subset X$
and a correspondence $\gamma \in \CH_d(D \times X)$ such that
$\gamma$ maps to $p$ via the inclusion $D \times X \r X \times X$.
Up to shrinking the open $U$ for which $p|_{U \times X}$ vanishes,
we can assume that $D$ has pure dimension $d-1$. Let $ Y \r D$ be an
alteration of $D$ and let $\sigma : Y \r D \hookrightarrow X$ be the
composite morphism. The induced map $\CH_d(Y \times X) \r \CH_d(D
\times X)$ is surjective and we have $p= (\sigma \times \id_X)_* f$,
where $f \in \CH_d(Y \times X)$ is a lift of $\gamma$. Then, by
\cite[16.1.1]{Fulton}, we have $(\sigma \times \id_X)_* f = f \circ
{}^t\Gamma_\sigma$. This yields a factorisation $p = f \circ g$,
where $f\in \CH_d(Y \times X)$ and $g = {}^t\Gamma_\sigma \in \CH^d(X
\times Y)$. Let's consider the correspondence $q:=g \circ f \circ g
\circ f = g \circ p \circ f \in \CH_{d-1}(Y \times Y)$. It is
straightforward to check that $q$ is an idempotent, and that $p
\circ f \circ q \circ g \circ p = p$ as well as $q \circ g \circ p
\circ f \circ q = q$. These last two equalities exactly mean that
$p \circ f \circ q$ seen as a morphism of Chow motives from
$(Y,q,1)$ to $(X,p,0)$ is an isomorphism with inverse $q \circ g
\circ p$.
\end{proof}
As noted by Sergey Gorchinskiy \cite{Gor}, this theorem admits the
following corollary
\begin{corollary} \label{effective-coro} Let $m$ and $n$ be positive
integers. Let $M=(X,p)$ be an effective Chow motive such that
$\CH_i(M_\Omega) = 0$ for $i\leq n-1$ and $\CH_j(M_\Omega^\vee(\dim
X)) = \CH_{j-\dim X}(M^\vee_\Omega) = 0$ for $j\leq m-1$. Then there
exists a smooth projective variety $Y$ of dimension $\dim X - m - n$
such that $M$ is isomorphic to a direct summand of $\h(Z)(n)$.
\end{corollary}
\begin{proof}
By Theorem \ref{effective} applied $n$ times, there is a smooth
projective variety $Y$ of dimension $\dim X - n$ and an idempotent
$q \in \CH_{\dim X -n}(Y\times Y)$ such that $M \simeq (Y,q,n)$. We
then have by duality $M^\vee(\dim X) \simeq (Y,{}^t q)$. Applying
$m$ times Theorem \ref{effective} gives a smooth projective variety
$Z$ of dimension $\dim Y - m = \dim X - n - m$ such that
$M^\vee(\dim X)$ is isomorphic to a direct summand of $\h(Z)(m)$.
Therefore, after dualizing, we see that $M$ is isomorphic to a
direct summand of $\h(Z)(n)$.
\end{proof}
\section{The Albanese motive and the Picard motive} \label{AP}
The results of the previous section show that it is convenient not
only to deal with smooth projective varieties but also with
idempotents. It may, however, be difficult to deal with idempotents
because these are usually not central. Here, we extend the construction
of Murre's Albanese projector to the case of Chow motives.\medskip
I thank Sergey Gorchinskiy \cite{Gor} for mentioning the following
basic lemma and the construction that ensues. The difference between
Lemma \ref{triangular} and Lemma \ref{linalg} is that Lemma
\ref{linalg} makes it possible to preserve self-duality when
orthonormalising a family of idempotents.
\begin{lemma}\label{triangular} Let $p$ be an idempotent endomorphism
of an object $A \oplus B$ in a Karoubi closed additive category. Let
$p_A$ denote the composition$$A \hookrightarrow A \oplus B
\stackrel{p}{\longrightarrow} A \oplus B \r A$$ and similarly for
$p_B$. Assume that $p$ is upper-triangular, that is, the composition
$$A \hookrightarrow A \oplus B \stackrel{p}{\longrightarrow}
A \oplus B \r B$$ vanishes. Then $p_A$ and $p_B$ are idempotents and
there is a canonical isomorphism
$$\im(p) \simeq
\im(p_A) \oplus \im(p_B).$$
\end{lemma}
\begin{proof} It is immediate that $p_A$ and $p_B$ are projectors. The
required isomorphism is given in the opposite direction by $p$.
\end{proof}
Given a smooth projective variety $X$ of dimension $d \geq 2$,
consider the decomposition constructed by Murre \cite{Murre}:
$$ \hspace{2cm} \h(X) =
\mathds{1} \oplus \h_1(X) \oplus M \oplus \h_{2d-1}(X) \oplus
\mathds{1}(d). \hspace{2cm} (*)$$ Since $\Hom(\mathds{1}(d),\h(X)) =
\Hom(\mathds{1}(d),\mathds{1}(d))$, we obtain
$$\Hom(\mathds{1}(d),\mathds{1}) = \Hom(\mathds{1}(d), \h_1(X)) =
\Hom(\mathds{1}(d),M) = \Hom(\mathds{1}(d), \h_{2d-1}(X)) =
0.$$
For any curve $C$, we have
$$\Hom(\h(C)(d-1),\h(X)) = \mathrm{Pic}(C \times X).$$
This implies that
$$\Hom(\h_{2d-1}(X),M) = 0.$$
By duality, we conclude that there are no morphisms going from the
right to the left in the decomposition of $\h(X)$ as above. By Lemma
\ref{triangular} applied several times, for any idempotent
endomorphism $p$ of $\h(X)$, we have a decomposition
$$\im(p) \simeq \im(p_0) \oplus \im(p_1) \oplus \im(p_M) \oplus
\im(p_{2d-1}) \oplus \im(p_{2d}),$$ where $\im(p)$ is a direct summand
in $\h(X)$ and where each direct summand appearing in the
decomposition of $\im(p)$ above is a direct summand of the
corresponding direct summand appearing in the decomposition $(*)$ of
$\h(X)$. The following proposition is then straightforward.
\begin{proposition} \label{albpic} Let $(X,p)$ be a motive. The
idempotents $p_0$, $p_1$, $p_{2d-1}$ and $p_{2d}$ constructed above
enjoy the following properties :
$\bullet$ $(X,p_0)$ is isomorphic to $\mathds{1}^{\oplus n}$ for
some $n$ and $H_*(X,p_0) = H_0(X,p)$.
$\bullet$ $(X,p_1)$ is isomorphic to a direct summand of the
$\h_1(C)$ for some curve $C$ and $H_*(X,p_1) = H_1(X,p)$.
$\bullet$ $(X,p_{2d-1})$ is isomorphic to a direct summand of the
$\h_1(C)(d-1)$ for some curve $C$ and $H_*(X,p_{2d-1}) =
H_{2d-1}(X,p)$.
$\bullet$ $(X,p_{2d})$ is isomorphic to $\mathds{1}({d})^{\oplus n}$
for some $n$ and $H_*(X,p_{2d}) = H_{2d}(X,p)$.
\end{proposition}
\begin{definition}
The idempotent $p_1$ is called the \emph{Albanese projector} and the
idempotent $p_{2d-1}$ is called the \emph{Picard projector}.
\end{definition}
\begin{remark}
The Albanese and Picard projectors are not unique.
\end{remark}
As an immediate corollary, we can extend Murre's theorem on surfaces
\cite{Murre} to direct summands of Chow motives of surfaces.
\begin{theorem} \label{CKSurface} Let $M = (S,p)$ be a Chow motive
where $S$ is a smooth projective surface. Then $M$ has a Murre
decomposition. If, moreover, $M$ is Kimura finite-dimensional
\cite{Kimura}, then $M$ satisfies Murre's conjecture (C).
\end{theorem}
\begin{proof}
The correspondences $p_0$, $p_1$, $p_3$ and $p_4$ of Proposition
\ref{albpic} together with $p_2 := p - \sum_{i \neq 2}p_i$ give a
Chow--K\"unneth decomposition for $M$. That such a decomposition
satisfies Murre's conjecture (B) is obvious. Murre's conjecture (D)
is possibly unclear only for one-cycles. Given a motive $N$, a
correspondence $\gamma \in \Hom(\h(S),N)$ that acts trivially on
$H^1(S)$ acts trivially on $\mathrm{Pic}^0_S = \CH^1(S)_\hom$. Since
$p_2$ and $p_3$ are the only projectors that act possibly
non-trivially on $\CH_1(S)$ and since $(p_3)_*\CH_1(M) \subseteq
\CH_1(M)_\hom$, we get conjecture (D) for one-cycles on $M$, see also
Proposition \ref{MurreC}. In the case when $M$ is Kimura
finite-dimensional, Murre's conjecture (C) for $M$ can be obtained
by applying \cite[Proposition 3.1]{Vial2}.
\end{proof}
\section{Self-dual Chow--K\"unneth decompositions}
\label{CK}
Let $X$ be a smooth projective variety of dimension $d$ over $k$. It
is proved in \cite[Theorem 4.2]{Vial3} that if the cohomology of $X$
in degree $\neq d$ is generated by the cohomology of curves, then $X$
admits a self-dual Chow--K\"unneth decomposition. Precisely, if
$H_i(X)=N^{\lfloor i/2 \rfloor}H_i(X)$ for all $i \neq d$, where $N$
is the coniveau filtration, then $X$ has a Chow--K\"unneth
decomposition. It follows from Theorem \ref{isogeny}, together with a
decomposition of the diagonal argument \`a la Bloch--Srinivas, that a
fourfold which is fibred by rationally connected threefolds over a
curve has a self-dual Chow--K\"unneth decomposition \cite[Corollary
4.7]{Vial3}. Del Angel and M\"uller-Stach \cite{dAMS} proved that
unirational threefolds have a Chow--K\"unneth decomposition. To do so,
they use Mori theory to reduce to the case of a conic fibration. In
this section, we generalise the aforementioned results by proving the
following:
\begin{theorem}\label{theoremCK}
Let $f:X \r S$ be a dominant morphism defined over a field $k$ from
a smooth projective variety $X$ to a smooth projective surface $S$
such that the general fibre of $f_\Omega$ has trivial Chow group of
zero-cycles. Suppose that $X$ has dimension $d \leq 4$. Then $X$
has a self-dual Chow--K\"unneth decomposition $\{p_i\}_{0 \leq i
\leq 2d}$.
\noindent Moreover, this decomposition can be chosen so as to
satisfy the following properties: \medskip
$\bullet$ $p_{0}$ factors through a point $P_0$, i.e. $(X,p_{0})$
is isomorphic to $\h(P_0)$.
$\bullet$ $p_{1}$ and $p_3$ factor through a curve, i.e. there
is a curve $C_0$ (resp. $C_1$) such that $(X,p_{1})$ (resp.
$(X,p_3)$) is a direct summand of $\h_1(C_0)$ (resp.
$\h_1(C_1)(1)$).
$\bullet$ $p_{2}$ factors through a surface, i.e. there is a
surface $S_0$ such that $(X,p_{2})$ is isomorphic to a direct
summand of $\h(S_0)$.
$\bullet$ If $d\leq 4$, $p_4$ factors through a surface, i.e. there
is a surface $S_1$ such that $(X,p_{4})$ is isomorphic to a direct
summand of $\h(S_1)(1)$.
\end{theorem}
\noindent In particular, this will give an alternate proof to del
Angel and M\"uller-Stach's result for conic fibrations over a surface.
\noindent We divide the proof into several steps.
\subsection{The projectors $\pi_0$, $\pi_1$ and $\pi_2^{tr}$}
\label{firstproj}
The surface $S$ has a Chow--K\"unneth decomposition \cite{Murre,
Scholl} $\{\pi_0^S, \pi_1^S, \pi_2^S, {}^t\pi_1^S, {}^t\pi_0^S\}$.
The motive $(S, \pi_2^S)$ admits a direct summand $(S, \pi_2^{tr,S})$
called its transcendental part, cf \cite{KMP}. The action of the
idempotent $\pi_2^{tr,S}$ on the homology of $S$ is the orthogonal
projector on the orthogonal complement for cup-product of the span of
the classes of algebraic one-cycles inside $H_2(S)$. In characteristic
zero, for Betti cohomology, $(\pi_2^{tr,S})_*H_*(S)$ is the sub-Hodge
structure of $H_2(S)$ generated by $H^{2,0}(S)=H^2(S,O_S)$ thanks to
the Lefschetz $(1,1)$-theorem. The idempotent $\pi_2^{tr,S}$ acts
trivially on $\CH_1(S_\Omega)$ and on $\CH_2(S_\Omega)$ so that
$\CH_*(S,\pi_2^{tr,S}) = \CH_0(S,\pi_2^{tr,S})$. \medskip
By Proposition \ref{surjective}, there is a correspondence $\Gamma \in
\CH^{d}(S \times X)$ such that $\Gamma_f \circ \Gamma = \Delta_S$, so
that the Chow motive of $S$ is a direct summand of the Chow motive of
$X$. We thus define mutually orthogonal idempotents $\pi_0 := \Gamma
\circ \pi_0^S \circ \Gamma_f$, $\pi_1 := \Gamma \circ \pi_1^S\circ
\Gamma_f$ and $\pi_2^{tr} := \Gamma \circ \pi_2^{tr, S} \circ
\Gamma_f$. Because the idempotents $\pi_0^S$, $\pi_1^S$ and
$\pi_2^{tr, S}$ in $\CH_2(S \times S)$ satisfy $(\pi_0^S)_*H_*(S) =
H_0(S)$, $(\pi_1^S)_*H_*(S) = H_1(S)$ and $(\pi_2^{tr,S})_*H_*(S)
\subset H_2(S)$, we see that $(\pi_0)_*H_*(X) \subset H_0(X)$,
$(\pi_1)_*H_*(X) \subset H_1(X)$ and $(\pi_2^{tr})_*H_*(X) \subset
H_2(X)$.\medskip
Since $\CH_0(S_\Omega) = (\pi_0^S + \pi_1^S + \pi_2^{tr,
S})_*\CH_0(S_\Omega)$ and since both $(\Gamma_f)_* : \CH_0(X_\Omega)
\r \CH_0(S_\Omega)$ and $\Gamma_* : \CH_0(S_\Omega) \r \CH_0(X_\Omega)$
are isomorphisms by Theorem \ref{isogeny}, we get
\begin{proposition} \label{vanishingchowS} $(\pi_0 + \pi_1 +
\pi_2^{tr})_*\CH_0(X_\Omega) = \CH_0(X_\Omega)$. \qed
\end{proposition}
This yields that the decomposition $\h(X) = (X,\pi_0) \oplus (X,\pi_1)
\oplus (X,\pi_2^{tr}) \oplus M$ satisfies $\CH_0(M_\Omega)=0$. Theorem
\ref{effective} gives a smooth projective variety $Y$ of dimension one
less than the dimension $d$ of $X$ together with an idempotent $q \in
\CH_{d-1}(Y \times Y)$ such that $M \simeq (Y,q,1)$.\medskip
By definition $H_*(Y,q,1) = H_{*-2}(Y,q) = q_*H_{*-2}(Y)$.
Consequently, we see that $H_0(M)=0$ and also that $H_1(M)=0$.
Therefore, $(\pi_0)_*H_*(X) = H_0(X)$ and $(\pi_1)_*H_*(X)= H_1(X)$.
\medskip
It is interesting to note that we can show that the $\pi_i$'s act as
the K\"unneth projectors on homology only after having determined
their action on Chow groups.
\subsection{Chow--K\"unneth decomposition for $\dim X=3$}
\label{CK3}
When $\dim X=3$, the Chow motive $\h(X)$ decomposes as $(X,\pi_0)
\oplus (X,\pi_1) \oplus (X,\pi_2^{tr}) \oplus M$ where $M$ is
isomorphic to a motive $(Y,q,1)$ with $\dim Y = 2$. In other words,
$\h(X)$ is isomorphic to a direct sum of direct summands of twisted
motives of surfaces. Theorem \ref{CKSurface} then says that $X$ has a
Murre decomposition. This is made more precise in \S\ref{dim3}.
\subsection{A first approach to splitting the motive of $X$ when $\dim
X \geq 3$} \label{dimsup3}
Ultimately, our goal is to define a self-dual Chow--K\"unneth
decomposition for $X$ with $\dim X \leq 4$. Let's thus study the
orthogonality relations between the idempotents $\pi_0$, $\pi_1$,
$\pi^{tr}_2$ and their transpose ${}^t\pi_0$, ${}^t\pi_1$,
${}^t\pi^{tr}_2$. We already know that $\pi_0$, $\pi_1$ and
$\pi^{tr}_2$ are mutually orthogonal. For dimension reasons (see also
\S\ref{orthoproj}), the only possible missing orthogonality relations
concern $\pi^{tr}_2 \circ {}^t\pi^{tr}_2$ and ${}^t\pi^{tr}_2 \circ
\pi^{tr}_2$. The crucial point here is that $\pi^{tr}_2 \circ
{}^t\pi^{tr}_2 = 0$ and that ${}^t\pi^{tr}_2 \circ \pi^{tr}_2$ acts
trivially on $\CH_*(X_\Omega)$. Let's prove these facts.\medskip
Proposition \ref{pi2vanish} gives two proofs that $\pi^{tr}_2 \circ
{}^t\pi^{tr}_2 = 0$. The first proof relies on Lemma \ref{1way} and is
particular to our geometric situation. The second proof relies on
Lemma \ref{2way}; it is more general and could be useful in other
situations.\medskip
On the one hand, we have
\begin{lemma} \label{1way} Let $f : X \r S$ be a dominant map between
two smooth projective varieties with $\dim X > \dim S$. Then
$\Gamma_f \circ {}^t\Gamma_f = 0$.
\end{lemma}
\begin{proof}
By definition, we have $\Gamma_f \circ {}^t\Gamma_f =
(p_{1,3})_*(p_{1,2}^*{}^t\Gamma_f \cap p_{2,3}^*\Gamma_f)$, where
$p_{i,j}$ denotes projection from $S \times X \times S$ to the
$(i,j)$-th factor. These projections are flat morphisms, therefore,
by flat pullback, we have $p_{1,2}^*{}^t\Gamma_f = [{}^t\Gamma_f
\times S]$ and $p_{2,3}^*\Gamma_f = [S \times \Gamma_f]$. It is easy
to see that the closed subschemes ${}^t\Gamma_f \times S$ and $S
\times \Gamma_f$ of $S \times X \times S$ intersect properly. Their
intersection is given by $\{(f(x),x,f(x)) : x \in X\} \subset S
\times X \times S$. This is a closed subset of dimension $\dim X$
and its image under the projection $p_{1,3}$ has dimension $\dim S$,
which is strictly less than $\dim X$ by assumption. The projection
$p_{1,3}$ is a proper map and hence, by proper pushforward, we get
that $(p_{1,3})_* [\{(f(x),x,f(x)) \in S \times X \times S : x \in X
\}] =0$.
\end{proof}
On the other hand, we have the following two lemmas, the first of
which will be used later on.
\begin{lemma} \label{trivialaction} Let $\gamma \in \CH^0(V \times
W)$ be a correspondence such that $\gamma_*$ acts trivially on
zero-cycles. Then $\gamma = 0$.
\end{lemma}
\begin{proof} We can assume that $V$ and $W$ are both connected. The
cycle $\gamma$ is equal to $a\cdot [V \times W]$ for some $a \in
\Q$. Let $z$ be a zero-cycle on $V$. Then $\gamma_*z = a \cdot \deg
z \cdot [W]$. This immediately implies $a=0$.
\end{proof}
\begin{lemma} \label{2way} Let $\gamma \in \CH^1(V \times W)$ be a
correspondence such that both $\gamma_*$ and $\gamma^*$ act
trivially on zero-cycles after base-change to an algebraically
closed field over $k$. Then $\gamma = 0$.
\end{lemma}
\begin{proof} Since base-change to a field extension induces an
injective map on Chow groups with rational coefficients, we may
assume that the base field $k$ is algebraically closed. We may also
assume that $V$ and $W$ are connected. We have
$$\mathrm{Pic}(V \times W) = \mathrm{Pic}(V) \times [W] \oplus [V]
\times \mathrm{Pic}( W) \oplus
\mathrm{Hom}(\mathrm{Alb}_V,\mathrm{Pic}^0_W)\otimes \Q.$$ Let
$\varphi \in \mathrm{Hom}(\mathrm{Alb}_V,\mathrm{Pic}^0_W)\otimes
\Q$ be the component of $\gamma$ under the above decomposition. By
assumption, $\gamma_*$ acts trivially on $\CH_0(V)$ and, hence, also
on $\CH_0(V)_\hom$. The albanese map $\CH_0(V)_\hom \r
\mathrm{Alb}_V(k)$ is surjective and it follows that $\varphi=0$.
The cycle $\gamma$ is thus equal to $D_1 \times [W] + [V] \times
D_2$ for some divisors $D_1 \in \CH^1(V)$ and $D_2 \in \CH^1(W)$. Let
$z$ be a zero-cycle on $V$. Then $\gamma_*z = \deg z \cdot D_2$.
This immediately implies that $D_2=0$. Likewise, if $z \in \CH_0(W)$,
then $\gamma^* z =0$ implies $D_1=0$. We have thus proved that
$\gamma=0$.
\end{proof}
\begin{proposition} \label{pi2vanish} $\pi_2^{tr} \circ {}^t\pi_2^{tr}
=0$. More generally, $\Hom(
(X,{}^t\pi_2^{tr}),(X,\pi_2^{tr}))=0$.
\end{proposition}
\begin{proof}
From Lemma \ref{1way} and from the very definition of $\pi_2^{tr}$,
it is immediate that $\pi_2^{tr} \circ {}^t\pi_2^{tr} =0$. Let now
$\alpha$ be a correspondence in $\CH_d(X \times X)$. The
correspondence $\pi_2 \circ \alpha \circ {}^t\pi_{2}^{tr}$ factors
through a correspondence $\gamma \circ \pi_2^{tr,S} \in \CH_d(S
\times S)$. If $d >4$, then the statement is clear. If $d=4$, we use
the fact that $\pi_2^{tr,S}$ sends zero-cycles on $S$ to zero-cycles
in the Albanese kernel of $S$. Hence, $\gamma\circ \pi_2^{tr,S}$
sends zero-cycles on $S$ to homologically trivial two-cycles on $S$.
In particular, $\gamma\circ \pi_2^{tr,S}$ acts trivially on
zero-cycles on $S$ and we can therefore apply Lemma
\ref{trivialaction}. Let's now assume that $d=3$ and let's give a
proof using Lemma \ref{2way} when the base field is a subfield of
$\C$. The reason is that we use Abel--Jacobi maps (although it is
almost certainly true that the Albanese variety and the Picard
variety enjoy the required functoriality properties over any base
field). The correspondence $\pi_2^{tr} \circ \alpha \circ
{}^t\pi_2^{tr} \in \CH_3(X \times X)$ factors through a
correspondence $\pi_2^{tr, S} \circ \gamma \circ {}^t\pi_2^{tr,S}
\in \CH^1(S \times S)$. In particular, by functoriality of the
Abel--Jacobi map, $\pi_2^{tr, S} \circ \gamma \circ
{}^t\pi_2^{tr,S}$ sends $0$-cycles on $S_\C$ to $1$-cycles on $S_\C$
which lie in the kernel of the Abel--Jacobi map. This last kernel
is trivial. Therefore, $\pi_2^{tr, S} \circ \gamma \circ
{}^t\pi_2^{tr,S}$ acts trivially on zero-cycles on $S_\Omega$.
Clearly the same holds for its transpose. Therefore, $\pi_2^{tr, S}
\circ \gamma \circ {}^t\pi_2^{tr,S} = 0$ and, hence, $\pi_2^{tr} \circ
\alpha \circ {}^t\pi_2^{tr} =0$.
\end{proof}
\begin{proposition} \label{trivialactionp2}
${}^t\pi_2^{tr} \circ \pi_2^{tr} $ acts trivially on $\CH_*(X_\Omega)$.
\end{proposition}
\begin{proof}
The correspondence ${}^t\pi_2^{tr} \circ \pi_2^{tr} $ factors
through a correspondence $\gamma\circ \pi_2^{tr,S} \in \CH_{4-d}(S
\times S)$. The proposition follows immediately since the idempotent
$ \pi_2^{tr,S}$ acts trivially on $\CH_1(X_\Omega)$ and on
$\CH_2(X_\Omega)$.
\end{proof}
Let's then define $p_0:=\pi_0$, $p_1:=\pi_1$ and
$p_2^{tr}:=(1-\frac{1}{2}{}^t\pi_2^{tr}) \circ \pi_2^{tr}$. It is
clear that these are idempotents and that $\{p_0, p_1, p_2^{tr},
{}^tp_2^{tr}, {}^tp_1, {}^tp_0 \}$ is a set of mutually orthogonal
idempotents in $\CH_d(X \times X)$. This yields a splitting $$\h(X) =
(X,p_0) \oplus (X, p_1) \oplus (X, p_2^{tr}) \oplus (X, {}^tp_2^{tr})
\oplus (X, {}^tp_1) \oplus (X, {}^tp_0) \oplus M.$$
\begin{proposition} \label{vanishingchow} $(p_0 + p_1 +
p_2^{tr})_*\CH_0(X_\Omega) = \CH_0(X_\Omega)$.
\end{proposition}
\begin{proof} By Proposition \ref{trivialactionp2} we see that
$(p_2^{tr})_*x = (\pi_2^{tr})_*x$ for all $X \in \CH_*(X_\Omega)$. We
can therefore conclude with Proposition \ref{vanishingchowS}.
\end{proof}
\begin{theorem} \label{finedec} There exists a smooth projective
variety $Z$ of dimension $d-2$ and an idempotent $q \in \CH_{d-2}(Z
\times Z)$ such that $$\h(X) = (X,p_0) \oplus (X, p_1) \oplus (X,
p_2^{tr}) \oplus (X, {}^tp_2^{tr}) \oplus (X, {}^tp_1) \oplus (X,
{}^tp_0) \oplus (Z,q,1).$$
\end{theorem}
\begin{proof} The theorem is a combination of Proposition
\ref{vanishingchow} and Corollary \ref{effective-coro}.
\end{proof}
\begin{theorem} If $d \leq 4$, then $X$ has a Murre decomposition.
\end{theorem}
\begin{proof} The theorem follows from Theorem \ref{finedec} and
Theorem \ref{CKSurface}.
\end{proof}
Let's write $\h(X) = (X,p) \oplus (Z,q,1)$, where $p = p_0+ p_1+
p_2^{tr}+{}^tp_2^{tr}+ {}^tp_1+ {}^tp_0$. Although $\h(X) =
\h(X)^\vee(d) = (X,p) \oplus (Z,{}^tq,1)$, it is not clear that
$(Z,q,1)$ is self-dual, i.e. isomorphic to $(Z,{}^tq,1)$. Thus we
need to refine the above construction.
\subsection{The projectors $\pi_2^{alg}$ and $\pi_3$} \label{proj3}
Until \S \ref{dim3}, the dimension $d$ of $X$ is supposed to be $\geq
4$. \medskip
Let's go back to the situation and notations of \S \ref{firstproj}.
Let $p := \Delta_X - (\pi_0 + \pi_1 + \pi_2^{tr})$. We have the
decomposition $\h(X) = (X,\pi_0) \oplus (X,\pi_1) \oplus
(X,\pi_2^{tr}) \oplus M$ with $M = (X,p)$ isomorphic to $(Y,q,1)$.
Choose an isomorphism $f : (Y,q,1) \r M$ and let $g : M \r (Y,q,1)$ be
its inverse. Let $q_0^Y$ and $q_1^Y$ be respectively the point
projector and the Albanese projector of Proposition \ref{albpic} for
$(Y,q,0)$. We define idempotents $\pi_2^{alg} := f \circ q_0^Y \circ
g$ and $\pi_3 : = f \circ q_1^Y \circ g$. \medskip
These two idempotents are orthogonal and are obviously orthogonal to
the idempotents $\pi_0$, $\pi_1$ and $\pi_2^{tr}$ previously defined.
Their action on cohomology is the expected one: we have $H_2(X) =
H_2(X,\pi_2^{tr}) \oplus H_2(M)$ but $H_2(M) = H_0(Y,q) =
H_0(Y,\pi_0^Y)$. Therefore $\pi_2 := \pi_2^{tr} + \pi_2^{alg}$ induces
the K\"unneth projector $H_*(X) \r H_2(X) \r H_*(X)$. We also have
$H_3(X) = H_3(M) = H_1(Y,q) = H_1(Y,\pi_1^Y)$ and hence
$(\pi_3)_*H_*(X) = H_3(X)$.
\subsection{The remaining projectors} \label{remainingproj}
We now define $\pi_{2d} := {}^t\pi_0$, $\pi_{2d-1} := {}^t\pi_1$,
$\pi_{2d-2} := {}^t\pi_2$ and $\pi_{2d-3} := {}^t\pi_3$. By Poincar\'e
duality, these idempotents satisfy $(\pi_i)_*H_*(X) = H_i(X)$.
\subsection{Orthonormalising the projectors} \label{orthoproj}
We have the following non-commutative Gram--Schmidt process \cite[Lemma
2.12]{Vial3}
\begin{lemma} \label{linalg} Let $V$ be a $\Q$-algebra and let $k$ and
$n$ be positive integers. Let $\pi_0, \ldots, \pi_n$ be idempotents
in $V$ such that $\pi_i \circ \pi_j = 0$ whenever $i -j < k$ and $i
\neq j$. Then the endomorphisms $$p_i := (1-\frac{1}{2}\pi_n) \circ
\cdots \circ (1-\frac{1}{2}\pi_{i+1}) \circ \pi_i \circ
(1-\frac{1}{2}\pi_{i-1}) \circ \cdots \circ (1-\frac{1}{2}\pi_0)$$
define idempotents such that $p_i \circ p_j = 0$ whenever $i -j <
k+1$ and $i \neq j$.
\end{lemma}
Let's state an orthonormalisation result which is valid in a general
setting and that we can apply to our particular case of interest.
\begin{theorem} \label{GS} Let $X$ be a smooth projective variety of
dimension $d$. Let $\pi_0, \ldots, \pi_n \in \CH_d(X \times X)$ be
idempotents such that $\pi_r \circ \pi_s =0$ for all $ r < s $. Then
applying $n$ times the Gram--Schmidt process of Lemma \ref{linalg}
gives mutually orthogonal idempotents $p_0, \ldots, p_n$ such that
$(X,p_r) \simeq (X,\pi_r)$ for all $r$. Furthermore, \medskip
$\bullet$ if there exists $r$ such that
$(\pi_r)_*H_*(X) = H_r(X)$, then $(p_r)_*H_*(X) =
H_r(X)$.
$\bullet$ if $\pi_r = {}^t\pi_{n-r}$ for all $r$, then $p_r =
{}^tp_{n-r}$ for all $r$;
$\bullet$ if there exists $r$ such that $\pi_s \circ \pi_r$ acts
trivially on $\CH_*(X)$ for all $s > r$, then $(p_r)_*\CH_*(X) =
(\pi_r)_*\CH_*(X)$ inside $\CH_*(X)$.
\end{theorem}
\begin{proof} The idempotents $\pi_0, \ldots, \pi_n$ satisfy the
assumptions of Lemma \ref{linalg} with $k=1$. Therefore, after
having run $n$ times the orthonormalisation process of Lemma
\ref{linalg}, we get mutually orthogonal idempotents. In order to
prove the theorem, it suffices to prove each statement after each
application of the orthonormalisation process. Given $r$, the
isomorphism $(X,p_r) \simeq (X,\pi_r)$ is simply given by the
correspondence $\pi_r \circ p_r$ ; its inverse is $p_r \circ \pi_r$
as can be readily checked.
If $(\pi_r)_*H_*(X) = H_r(X)$, then the image of $\pi_r$ in $H_d(X
\times X) \simeq \Hom(H_*(X),H_*(X))$ is central. Therefore, if
$\pi_r \circ \pi_s =0$ for all $ r < s $, then the image of $\pi_s
\circ \pi_r$ in $H_d(X \times X)$ is trivial for all $s > r$. It is
then straightforward to conclude that $(p_r)_*H_*(X) = H_r(X)$.
If $\pi_r = {}^t\pi_{n-r}$ for all $r$, then it is straightforward
to check from the formula of Lemma \ref{linalg} that $p_r =
{}^tp_{n-r}$ for all $r$
Let's fix $r$. Given the isomorphism $(X,p_r) \simeq (X,\pi_r)$, it
is very tempting to conclude that $(p_r)_*\CH_*(X_\Omega) =
(\pi_r)_*\CH_*(X_\Omega)$ in $\CH_*(X_\Omega)$. However, this appears
not to be obvious at all and a careful analysis of the
non-commutative Gram--Schmidt process needs to be carried on. By
examining the formula defining the idempotent $p_r$, together with
the assumption that $\pi_s \circ \pi_r$ acts trivially on $\CH_*(X)$
for all $s > r$, we see that, for $x \in \CH_*(X_\Omega)$, we have
$(p_r)_*x = (\pi_r)_*x \in \CH_*(X_\Omega)$.
\end{proof}
We wish to apply Theorem \ref{GS} to the set of idempotents
$\{\pi_0,\pi_1,\pi_2,\pi_3, \pi_{2d-3}, \pi_{2d-2}, \pi_{2d-1},
\pi_{2d}\}$. In order to do so, we have to show that $\pi_i \circ
\pi_j = 0$ whenever $i < j$. We already know that $\pi_0$, $\pi_1$,
$\pi_2$ and $\pi_3$ are mutually orthogonal. Let's prove the missing
orthogonality relations. First we have :
\begin{itemize}
\item $\pi_0 \circ {}^t\pi_{0} = \pi_0 \circ {}^t\pi_{1} = \pi_0 \circ
{}^t\pi_{2} = \pi_0 \circ {}^t\pi_{3} = 0$.
\item $\pi_1 \circ {}^t\pi_{1} = \pi_1 \circ
{}^t\pi_{2} = \pi_1 \circ {}^t\pi_{3} = 0$.
\item $\pi_2 \circ {}^t\pi_{2}^{alg} = 0$ and hence $\pi_2 \circ
{}^t\pi_2 = 0$ thanks to Proposition \ref{pi2vanish}.
\end{itemize}
These relations are obvious : one uses a dimension argument as well
as the fact that $\pi_0$ (resp. $\pi_1$, $\pi_2^{tr}$,
$\pi_2^{alg}$, $\pi_3$) factors through a variety $P_0$ (resp.
$C_0$, $S$, $P_1$, $C_1$) of dimension $0$ (resp. $1$, $2$, $0$,
$1$). For instance, $\pi_1 \circ {}^t\pi_{3}$ factors through a
correspondence in $\CH_{d-1}(C_1 \times C_0)$. If $d \geq 4$, then
this last group is trivial.\medskip
Using the same arguments, the following orthogonality relations can
be further proved. These relations are not necessary to run the
non-commutative Gram--Schmidt process but are essential to the proof
of Proposition \ref{vanishingchow2}.
\begin{itemize}
\item $ {}^t\pi_0 \circ \pi_{0} = {}^t\pi_0 \circ \pi_{1} =
{}^t\pi_0 \circ \pi_{2} = {}^t\pi_0 \circ \pi_{3} = 0$.
\item $ {}^t\pi_1 \circ \pi_{1} = {}^t\pi_1 \circ \pi_{2} =
{}^t\pi_1 \circ \pi_{3} = 0$.
\item $ {}^t\pi_2 \circ \pi_{2}^{alg} = 0$.
\item $ {}^t\pi_3 \circ \pi_{2}^{alg} = 0$.
\end{itemize}
Secondly, the remaining orthogonality relations needed to run the
non-commutative Gram--Schmidt process follow from Lemma
\ref{trivialaction}.
\begin{itemize}
\item $\pi_2 \circ {}^t\pi_{3} = 0$. The correspondence $\pi_2 \circ
{}^t\pi_{3}$ factors through a correspondence $\gamma \in
\CH_{d-1}(C_1 \times S_0)$, where $S_0$ is a surface, that sends
zero-cycles to homologically trivial cycles on $S_0$. Again, if
$d>4$, then the result is trivial. If $d=4$, then we conclude by
Lemma \ref{trivialaction}.
\item $\pi_3 \circ {}^t\pi_{3} = 0$. The correspondence $\pi_3 \circ
{}^t\pi_{3}$ factors through a correspondence $\gamma \in
\CH_{d-2}(C_1 \times C_1)$ that sends zero-cycles to homologically
trivial cycles on $C$. Again, if $d>4$, then the result is
trivial. If $d=4$, then we conclude by Lemma \ref{trivialaction}.
\end{itemize}
We are now in a position to apply Theorem \ref{GS} to obtain a set of
mutually orthogonal idempotents $\{p_0,p_1,p_2,p_3, p_{2d-3},
p_{2d-2}, p_{2d-1}, p_{2d}\}$ such that $p_{2d-i}={}^tp_i$ which
induce the expected K\"unneth projectors modulo homological
equivalence.
\begin{remark} \label{missingorth} It follows from the above
discussion that the only possible missing orthogonality relations
among the idempotents $\pi_0,\pi_1,\pi_2^{alg},\pi_2^{tr}, \pi_3$
and their transpose are the following.
\begin{itemize}
\item ${}^t\pi_3 \circ \pi_2^{tr}$.
\item ${}^t\pi_2^{tr} \circ \pi_2^{tr}$.
\item ${}^t\pi_3 \circ \pi_3$.
\end{itemize}
It can then be checked that it is actually enough to run the
non-commutative Gram--Schmidt process only once on the set of
idempotents $\{\pi_0,\pi_1,\pi_2,\pi_3, \pi_{2d-3}, \pi_{2d-2},
\pi_{2d-1}, \pi_{2d}\}$ to obtain a set of mutually orthogonal
idempotents. We can therefore describe the $p_i$'s in terms of the
$\pi_i$'s by not too complicated explicit formulas. Such formulas
may then be used, for instance, to give a quicker proof of the motivic
Lefschetz conjecture for $X$. However, we describe a method that
might be useful in other situations where the Gram--Schmidt process
needs to be run several times.
\end{remark}
The following proposition is fundamental to proving Proposition
\ref{vanishingchow2} and, hence, to proving Murre's conjectures for
$X$.
\begin{proposition} \label{trivialactionchow} Let $p$ and $q$ be any
two distinct idempotents among the idempotents $\pi_0$, $\pi_1$,
$\pi_2$, $\pi_3$, $\pi_{2d-3}$, $\pi_{2d-2}$, $\pi_{2d-1}$ and
$\pi_{2d}$. Then $p \circ q$ acts trivially on $\CH_l(X_\Omega)$ for
all $l$.
\end{proposition}
\begin{proof}
From remark \ref{missingorth} we only need to prove that ${}^t\pi_3
\circ \pi_2^{tr}$, ${}^t\pi_2^{tr} \circ \pi_2^{tr}$ and ${}^t\pi_3
\circ \pi_3$ act trivially on $\CH_*(X_\Omega)$. In the first case,
${}^t\pi_3 \circ \pi_2^{tr}$ factors through a correspondence
$\gamma \circ \pi_2^{tr,S} \in \CH_0(S \times C_1)$ for some curve
$C_1$ and it therefore acts trivially on $\CH_*(X_\Omega)$ because $
\pi_2^{tr,S}$ only acts possibly non-trivially on $\CH_0(S_\Omega)$.
In the second case, ${}^t\pi_2^{tr} \circ \pi_2^{tr}$ factors
through a correspondence $\gamma \circ \pi_2^{tr,S} \in \CH_0(S
\times S)$ and we conclude in the same way. In the last case,
${}^t\pi_3 \circ \pi_3$ factors through a correspondence $\gamma
\circ \pi_1^{C_1} \in \CH_0(C_1 \times C_1)$ which also acts
trivially on $\CH_*((C_1)_\Omega)$ because $\pi_1^{C_1}$ acts
trivially on $\CH_1((C_1)_\Omega)$.
\end{proof}
\begin{proposition} \label{vanishingchow2}
$(p_0 + p_1 + p_2)_*\CH_0(X_\Omega) = \CH_0(X_\Omega)$.
\end{proposition}
\begin{proof}
Proposition \ref{trivialactionchow} and Theorem \ref{GS} imply that
$(\pi_i)_*x = (p_i)_*x$ for $i = 0, 1$ or $2$ and for all $x \in
\CH_0(X_\Omega)$. By Proposition \ref{vanishingchowS}, this yields
$(p_0+p_1 + p_2)_*\CH_0(X_\Omega) = \CH_0(X_\Omega)$ as claimed.
\end{proof}
Finally, when $d=4$, we define $p_4:=\Delta_X - \sum_{i \neq 4} p_i$.
The set $\{p_i\}_{0 \leq i \leq 8}$ is then a self-dual Chow--K\"unneth
decomposition for $X$. Moreover, $p_4$ has the following property.
\begin{proposition}
$(X,p_4)$ is isomorphic to a direct summand of $\h(S_1)(1)$ for some
smooth projective surface $S_1$.
\end{proposition}
\begin{proof}
Proposition \ref{vanishingchow2} implies that
$(p_4)_*\CH_0(X_\Omega)=0$. Also we know that $p_4 = {}^tp_4$.
Therefore, the result follows immediately from Corollary
\ref{effective-coro}.
\end{proof}
\subsection{Back to the case $\dim X = 3$} \label{dim3}
Let's now consider the case of a conic fibration over a surface. In
section \ref{CK3}, we already gave a quick argument showing that $X$
has a Chow--K\"unneth decomposition. As in the case $\dim X =4$, we
want to show that a Chow--K\"unneth decomposition for $X$ can be chosen
to be self-dual, a result which is not shown in \cite{dAMS}. In order
to prove Murre's conjectures for such a decomposition (which will be
done in section \ref{Murreconj}), we also want to show that the middle
idempotent factors through a curve. \medskip
For this purpose, we define $\pi_0$, $\pi_1$, $\pi_2^{tr}$,
$\pi_2^{alg}$ and $\pi_2 := \pi_2^{tr} + \pi_2^{alg}$ the same way we
did in sections \ref{firstproj} and \ref{proj3}. The only difference
with section \ref{proj3} is that we don't define an idempotent
$\pi_3$. We then define $\pi_6 = {}^t\pi_0$, $\pi_5 = {}^t\pi_1$ and
$\pi_4 = {}^t\pi_2$. As in sections \ref{firstproj} and \ref{proj3},
it is easy to see that these do define the K\"unneth projectors in
homology.
As before, we have $\pi_i \circ \pi_j = 0$ for all $i<j$ not equal to
$3$. These relations make it possible to run the non-commutative
Gram--Schmidt process and to get mutually orthogonal
idempotents $p_0, p_1, p_2, p_4, p_5, p_6$ such that
$(p_i)_*H_*(X)=H_i(X)$ and $p_{6-i} = {}^tp_i$ for all $i \neq 3$.
Setting $p_3 := \Delta_X - \sum_{i\neq 3} p_i$, we thus get a
self-dual Chow--K\"unneth decomposition for $X$. Again, as before, we
have that $\pi_j \circ \pi_i$ acts trivially on $\CH_*(X_\Omega)$ for
all $j>i$ not equal to $3$. The middle idempotent $p_3$ has thus the
following property.
\begin{proposition} \label{middle-curve}
There exists a curve $C_1$ such that $(X,p_3)$ is isomorphic to a
direct summand of $\h_1(C_1)(1)$.
\end{proposition}
\begin{proof} As in the proof of Proposition \ref{vanishingchow2}, we
have $(p_0 + p_1 + p_2)_*\CH_0(X_\Omega) = \CH_0(X_\Omega)$. Therefore,
$\CH_0(X_\Omega,p_3)=0$. But $p_3 = {}^tp_3$, so that
$\CH_0(X_\Omega,{}^tp_3)=0$ too. It follows from Corollary
\ref{effective-coro} that there exists a curve $C_1$ such that
$(X,p_3)$ is isomorphic to a direct summand of $\h(C_1)(1)$. The
fact that the motive $(X,p_3)$ is pure of weight $3$ makes it
possible to conclude.
\end{proof}
\section{The motivic Lefschetz conjecture for $X$} \label{motLef}
Let $X$ be a smooth projective variety of dimension $d$ over a field
$k$. Let $i \leq d$ and let $\iota : H \r X$ be a smooth linear
section of dimension $i$ and let $L := (\iota,\id_X)_*\Gamma_\iota =
\Gamma_{\iota} \circ {}^t\Gamma_\iota \in \CH_{i}(X \times X)$. The
correspondence $L$ acts on cohomology or Chow groups as intersecting
$d-i$ times by a smooth hyperplane section of $X$. The variety $X$ is
said to satisfy the motivic Lefschetz conjecture in degree $i$ if
there exist mutually orthogonal idempotents $\pi_i$ and $\pi_{2d-i}$
such that $H_*(X,\pi_i)=H_i(X)$ and $H_*(X,\pi_{2d-i})=H_{2d-i}(X)$
and such that the induced map $$L : (X,\pi_{2d-i}) \r (X,\pi_i,d-i)$$
is an isomorphism of Chow motives. The variety $X$ is said to satisfy
the \emph{motivic Lefschetz conjecture} if it satisfies the motivic
Lefschetz conjecture in all degrees $<d$. Note that if $X$ satisfies
the motivic Lefschetz conjecture in degree $i$ then $X$ satisfies the
Lefschetz standard conjecture in degree $i$, i.e. there exists a
correspondence $\Gamma \in \CH^{i}(X \times X)$ such that $\Gamma_* :
H_i(X) \r H_{2d-i}(X)$ is the inverse to $L : H_{2d-i}(X) \r H_i(X)$.
The motivic Lefschetz conjecture for $X$ follows from a combination of
the Lefschetz standard conjecture for $X$ and of Kimura's finite
dimensionality conjecture for $X$ ; it is thus expected to hold for
all smooth projective varieties.
\begin{proposition} \label{point} Let $P$ be a zero-dimensional
variety over $k$. Let $p \in \CH_d(X \times X)$ be an idempotent
such that $(X,p)$ is isomorphic to $\h(P)(i)$ for some integer $i$
satisfying $2i \leq d$. If the induced map $L : H_{2d-2i}(X,{}^tp)
\r H_{2i}(X,p)$ is an isomorphism, then $L : (X,{}^tp) \r
(X,p,d-2i)$ is an isomorphism of Chow motives.
\end{proposition}
\begin{proof}
There exist correspondences $f \in \Hom(\h(P)(i),(X,p))$ and $g \in
\Hom((X,p),\h(P)(i))$ such that $g\circ f = \id_{\h(P)(i)}$ and $f
\circ g = p$. The correspondence $g \circ L \circ {}^tg \in
\End(\h(P))$ induces an automorphism of $H_{0}(P)$ and, hence, is
itself an automorphism. Therefore, it admits an inverse $\alpha \in
\End(\h(P))$. It is now straightforward to check that ${}^tp
\circ {}^tg \circ \alpha \circ g \circ p$ is the inverse of $p \circ
L \circ {}^tp$.
\end{proof}
\begin{proposition} \label{curve} Let $J$ be an abelian variety over
$k$. Let $p \in \CH_d(X \times X)$ be an idempotent such that
$(X,p)$ is isomorphic to $\h_1(J)(i)$ for some integer $i$
satisfying $2i+1 \leq d$ and such that $p$ is orthogonal to ${}^tp$
(this last condition is automatically satisfied if $2i+1 <d-1$). If
the induced map $L : H_{2d-2i-1}(X,{}^tp) \r H_{2i+1}(X,p)$ is an
isomorphism, then $L : (X,{}^tp) \r (X,p,d-2i-1)$ is an isomorphism
of Chow motives.
\end{proposition}
\begin{proof}
There exist correspondences $f \in \Hom(\h_1(J)(i),(X,p))$ and $g
\in \Hom((X,p),\h_1(J)(i))$ such that $g\circ f = \id_{\h_1(J)(i)}$
and $f \circ g = p$. The correspondence $g \circ L \circ {}^tg \in
\End(\h_1(J))$ induces an automorphism of $H_{1}(J)$ and, hence, is
itself an automorphism (indeed by \cite[Prop. 4.5]{Scholl} we have
$\End(\h_1(J)) = \End_k(J) \otimes \Q$ and it is well-known that a
map between abelian varieties which induces an isomorphism in degree
one homology must be an isogeny). Therefore, it admits an inverse
$\alpha \in \End(\h_1(J))$. It is now straightforward to check that
${}^tp \circ {}^tg \circ \alpha \circ g \circ p$ is the inverse of
$p \circ L \circ {}^tp$.
\end{proof}
As already proved by Scholl \cite{Scholl}, every smooth projective
variety satisfies the motivic Lefschetz conjecture in degrees $\leq
1$.
\begin{theorem}
Let $f : X \r S$ be a dominant morphism defined over a field $k$
from a smooth projective variety $X$ to a smooth projective surface
$S$ such that the general fibre of $f_\Omega$ has trivial Chow group
of zero-cycles. Then $X$ satisfies the motivic Lefschetz
conjecture in degrees $\leq 3$. In particular, if $X$ has dimension
$\leq 4$, then $X$ satisfies the motivic Lefschetz conjecture and
hence the Lefschetz standard conjecture.
\end{theorem}
\begin{proof}
By Theorem \ref{theoremCK}, $p_0$ factors through a
point, and $p_1$ and $p_3$ factor through the $\h_1$ of a curve. The
hard Lefschetz theorem says that the map $H_{2d-i}(X) \r H_i(X)$
induced by intersecting $d-i$ times with a smooth hyperplane section
is an isomorphism. Therefore, the two propositions above give the
motivic Lefschetz conjecture in degrees $0$, $1$ and $3$ for
$X$.\medskip
Let $\pi_2^{tr}$ be the idempotent of section 4.1. Let's prove that
$L : (X,{}^t\pi_2^{tr},0) \r (X, \pi_2^{tr}, d-2)$ is an isomorphism
of Chow motives. Because $\iota : H \r X$ is a linear section of $X$
of dimension $2$, Proposition \ref{surjective} gives a non-zero
integer $m$ such that $\Gamma_f \circ L \circ {}^t\Gamma_f = m \cdot
\Delta_S$. It is then straightforward to check that $\frac{1}{m}
\cdot {}^t\pi_2^{tr} \circ {}^t\Gamma_f \circ \Gamma_f \circ
\pi_2^{tr}$ is the inverse of $ \pi_2^{tr} \circ L \circ {}^t
\pi_2^{tr}$.
Let $\pi_2^{alg}$ be the idempotent of section 4.3. Because $L :
(X,{}^t\pi_2^{tr},0) \r (X, \pi_2^{tr}, d-2)$ is an isomorphism and
because $L_* : H_{2d-2}(X) \r H_2(X)$ is an isomorphism by the hard
Lefschetz theorem, we see that $L$ induces an isomorphism $L :
H_{2d-2}(X,{}^t\pi_2^{alg}) \r H_2(X,\pi_2^{alg})$. Proposition
\ref{point} then shows that $ \pi_2^{alg} \circ L \circ {}^t
\pi_2^{alg} \in \Hom((X,{}^t\pi_2^{alg}),(X,\pi_2^{alg},d-2))$ is an
isomorphism.
We have thus showed that $ \pi_2 \circ L \circ {}^t \pi_2 \in
\Hom((X,{}^t\pi_2),(X,\pi_2,d-2))$ is an isomorphism. Since, by
Theorem \ref{GS}, we know that $(X,p_2,d-2)\simeq (X,\pi_2,d-2)$ and
$(X,{}^tp_2) \simeq (X,{}^t\pi_2)$, we get that $(X,p_2,d-2)$ is
isomorphic to $(X,{}^tp_2)$. However, the isomorphism is induced by
$p_2 \circ \pi_2 \circ L \circ {}^t \pi_2 \circ {}^t p_2 $ which is
not quite the isomorphism we were aiming at. \medskip
By Remark \ref{missingorth}, it can be checked that in our
particular setting we have $p_2 \circ \pi_2 = p_2$ so that $ p_2
\circ L \circ {}^t p_2$ is an isomorphism with inverse $\frac{1}{m}
\cdot {}^tp_2 \circ {}^t\pi_2 \circ {}^t\Gamma_f \circ \Gamma_f
\circ \pi_2 \circ p_2$.
Let's however give another proof that $ p_2 \circ L \circ {}^t p_2$
is an isomorphism that might be useful in other situations. We can
conclude that $ p_2 \circ L \circ {}^t p_2$ is an isomorphism if we
can show that it is equal to $ p_2 \circ \pi_2 \circ L \circ {}^t
\pi_2 \circ {}^t p_2$. For this purpose, after examining the formula
of Lemma \ref{linalg} defining $p_2$, it is enough to check that, for
all correspondences $\alpha \in \CH_{2}(X \times X)$, we have $\pi_r
\circ \alpha \circ {}^t \pi_2 = 0$ for $r = 0,1$. This is recorded
in the lemma below.
\end{proof}
\begin{lemma}
$\Hom ((X,{}^t\pi_2),(X,\pi_r,d-2)) = 0$ for $r=0$ or $1$.
\end{lemma}
\begin{proof}
When $r=0$, $\pi_0 \circ \alpha \circ {}^t \pi_2^{alg}$ factors
through a correspondence $\gamma \in \CH_1(P_1 \times P_0)$ for some
zero-dimensional $P'$ and is thus zero for dimension reasons and
$\pi_0 \circ \alpha \circ {}^t \pi_2^{tr}$ factors through a
correspondence $\gamma \in \CH^0(S \times P_0)$ with $\gamma^* z =0$
for any $z \in \CH_0(P)$. Lemma \ref{trivialaction} then shows that
$\gamma =0$ and hence $\pi_0 \circ \alpha \circ {}^t \pi_2^{tr} =
0$. When $r=1$, on the one hand, we have that $\pi_1 \circ \alpha
\circ {}^t\pi_2^{alg}$ factors through a correspondence $\gamma \in
\CH^0(P_1 \times C_0)$ with $\gamma^* z =0$ for any zero-cycle $z$ on
$C_0$. Lemma \ref{trivialaction} then shows that $\gamma =0$ and
hence $\pi_1 \circ \alpha \circ {}^t\pi_2^{alg} = 0$. On the other
hand, $\pi_1 \circ \alpha \circ {}^t\pi_2^{tr}$ factors through a
correspondence $\gamma \in \CH^1(S \times C_0)$ with $\gamma^* z =0$
for any zero-cycle $z$ on $(C_0)_\Omega$ and $\gamma_* z'$ for any
zero-cycle $z'$ on $S_\Omega$ by functoriality of the Abel--Jacobi
map. Therefore thanks to Lemma \ref{2way}, we get $\gamma=0$ and
hence $\pi_1 \circ \alpha \circ {}^t\pi_2^{tr}=0$.
\end{proof}
\begin{remark}
The results of this section actually show that, for $X$ as in the
theorem above and for the idempotents $p_i$ constructed in \S
\ref{CK}, the map $L : (X,p_{2d-i}) \r (X,p_i,d-i)$ is an isomorphism
for $i \leq 3$ for any choice of a smooth linear section $\iota : H
\hookrightarrow X$ of dimension $i$.
\end{remark}
\section{Murre's conjectures for $X$} \label{Murreconj}
As shown by Jannsen \cite{Jannsen}, Murre's conjectures \cite{Murre1}
are equivalent to Bloch and Beilinson's. Let's recall them : \medskip
(A) $X$ has a Chow--K\"unneth decomposition $\{\pi_0, \ldots,
\pi_{2d}\}$ : There exist mutually orthogonal idempotents $\pi_0,
\ldots, \pi_{2d} \in \CH_d(X \times X)$ adding to the identity such
that $(\pi_i)_*H_*(X)=H_i(X)$ for all $i$.
(B) $\pi_0, \ldots, \pi_{2l-1},\pi_{d+l+1}, \ldots, \pi_{2d}$ act
trivially on $\CH_l(X)$ for all $l$.
(C) $F^i\CH_l(X) := \ker(\pi_{2l}) \cap \ldots \cap \ker(\pi_{2l+i-1})$
doesn't depend on the choice of the $\pi_j$'s. Here the $\pi_j$'s are
acting on $\CH_l(X)$.
(D) $F^1\CH_l(X) = \CH_l(X)_\hom$. \\
Before we consider Murre's conjectures for $X$ as in Theorem
\ref{mainth}, let's consider the following situation. Let $\Pi \in
\CH_d(X\times X) = \End(\h(X))$ be an idempotent which factors as
$$\h(X) \stackrel{g}{\longrightarrow} \h(Y)
\stackrel{f}{\longrightarrow} \h(X)$$ where $Y$ is a smooth projective
variety of dimension $\leq l+1$. (Actually, up to replacing $Y$ with
$Y \times \P^{l+1-\dim Y}$, we can assume $\dim Y =l+1$.) The
arguments in the proof of Propositions \ref{MurreB} and \ref{MurreC}
below are essentially contained in \cite{Vial2} and \cite{Vial3}.
\begin{proposition} \label{MurreB} Let $\Pi$ be as above.
$\bullet$ If $\Pi_*H_{2l+1}(X) = 0$, then $\Pi$ acts trivially on
$\CH_l(X)_\hom$.
$\bullet$ If, moreover, $\Pi_*H_{2l}(X) = 0$, then $\Pi$ acts
trivially on $\CH_l(X)$.
\end{proposition}
\begin{proof} Because $\Pi$ is an idempotent, we see that $g \circ f$
acts trivially on $H^1(Y)$. Therefore, $g \circ f$ acts trivially on
$\CH^1(Y)_\hom$. Thus, $\Pi$ acts trivially on $\CH_l(X)_\hom$. If,
moreover, $\Pi_*H_{2i}(X) = 0$, then $g \circ f$ acts trivially on
$\CH^1(Y)$. Thus, $\Pi$ acts trivially on $\CH_l(X)$.
\end{proof}
\begin{proposition} \label{MurreC} Let $\Pi$ be as above.
$\bullet$ If $\Pi_*H_{*}(X) = H_{2l}(X)$, then $\ker \big( \Pi_* :
\CH_l(X) \r \CH_l(X)\big) = \CH_l(X)_\hom$.
$\bullet$ Assume that $k \subseteq \C$. If $\Pi_*H_{*}(X) =
H_{2l+1}(X)$, then $\ker \big( \Pi_* : \CH_l(X)_\hom \r \CH_l(X)_\hom
\big) = \ker \big( AJ_l : \CH_l(X)_\hom \r J_l(X) \otimes \Q \big)$.
Here $AJ_l$ is the Abel--Jacobi map.
\end{proposition}
\begin{proof} In the first case, the inclusion $\subseteq$ follows
immediately from the functoriality of the cycle class map with
respect to the action of correspondences. The reverse inclusion
$\supseteq$ follows from the first point of Proposition
\ref{MurreB}.
In the second case, we consider the Abel--Jacobi map $AJ_l :
\CH_l(X)_\hom \r J_l(X) \otimes \Q$ instead of the cycle class map
$\CH_l(X) \r H_{2l}(X)$. The Abel--Jacobi map is functorial with
respect to the action of correspondences and, if a correspondence
$\alpha \in \End(\h(X))$ induces the identity on $H_{2l+1}(X)$, then
$\alpha_*$ induces the identity on $J_l(X) \otimes \Q$. This yields
a commutative diagram
\begin{center} $ \xymatrix{ \CH_l(X)_\hom \ar[d]^{AJ_l} \ar[r]^{g_*}
& \CH^1(Y)_\hom \ar[d] \ar[r]^{f_*} & \CH_l(X)_\hom
\ar[d]^{AJ_l} \\
J_{l}(X)(\C)\otimes \Q \ar[r] & \mathrm{Pic}^0_{Y}(\C)\otimes \Q
\ar[r] & J_{l}(X)(\C)\otimes \Q.}$ \end{center} where the
composite of the two bottom arrows is the identity and where the
middle vertical arrow is injective. It is then straightforward to
conclude by a simple diagram chase.
\end{proof}
\begin{definition} \label{special} A smooth projective variety $X$ of
dimension $d$ is said to have a \emph{special} Chow--K\"unneth
decomposition $\{\pi_i\}_{0 \leq i \leq 2d}$ if, for all $i$,
$\bullet$ $\pi_{2i}$ factors through a surface, i.e. there is a
surface $S_i$ such that $(X,\pi_{2i})$ is a direct summand of
$\h(S_i)(i-1)$.
$\bullet$ $\pi_{2i+1}$ factors through a curve, i.e. there is a
curve $C_i$ such that $(X,\pi_{2i+1})$ is a direct summand of
$\h_1(C_i)(i)$.
\end{definition}
\begin{proposition} \label{specialMurre} Let $X$ be a smooth
projective variety that has a special Chow--K\"unneth decomposition.
Then homological and algebraic equivalence agree on $X$, and $X$
satisfies Murre's conjectures (A), (B) and (D). Moreover, if $ k
\subseteq \C$, then the filtration $F$ on $\CH_l(X)$ does not depend
on the choice of a special Chow--K\"unneth decomposition for $X$.
\end{proposition}
\begin{proof} That homological and algebraic equivalence agree on $X$
follows from the well-known fact that they agree on zero-cycles and
on codimension-one cycles. That $X$ satisfies Murre's conjectures
(B) is obvious and that $X$ satisfies (D) follows from the first
point of Proposition \ref{MurreC}. That the induced filtration on
the Chow groups of $X$ is independent of the choice of a special
Chow--K\"unneth decomposition for $X$ is contained in Proposition
\ref{MurreC}.
\end{proof}
Since the Chow--K\"unneth decomposition of $X$ as in Theorem
\ref{theoremCK} is a special Chow--K\"unneth decomposition, we can
state the following.
\begin{theorem} \label{Murre-theorem} Let $f : X \r S$ be a dominant
morphism defined over a field $k$ between a smooth projective
variety $X$ of dimension $\leq 4$ and a smooth projective surface
$S$ such that the general fibre of $f_\Omega$ has trivial Chow group
of zero-cycles. Then $X$ has a special Chow--K\"unneth decomposition
which is self-dual and which satisfies Murre's conjectures (B) and
(D) and, if $ k \subseteq \C$, then the induced filtration $F$
on $\CH_l(X)$ does not depend on the choice of a special
Chow--K\"unneth decomposition for $X$. Finally, whichever the
characteristic of $k$ is, if $X$ is Kimura finite-dimensional, then
$F$ does not depend on the choice of a Chow--K\"unneth decomposition
for $X$.
\end{theorem}
\begin{proof}
Theorem \ref{theoremCK} says that $X$ has a self-dual
Chow--K\"unneth decomposition $\{p_i : 0 \leq i \leq 2d\}$ which is
special. We may then conclude with Proposition \ref{specialMurre}
that $X$ satisfies Murre's conjectures (B) and (D) and that, if
$\mathrm{char} \ k=0$, then the induced filtration $F$ on $\CH_l(X)$
does not depend on the choice of a special Chow--K\"unneth
decomposition for $X$. When $X$ is Kimura finite-dimensional,
Murre's conjecture (C) follows from applying \cite[Proposition
3.1]{Vial2} to $X$ endowed with the Chow--K\"unneth decomposition
given by the $p_i$'s.
\end{proof}
\section{Murre's conjectures for $X \times C$} \label{Murreconj2}
Let $f : X \r S$ be a dominant morphism from a smooth projective
fourfold to a smooth projective surface such that the general fibre of
$f_\Omega$ has trivial Chow group of zero-cycles. Consider the
self-dual Chow--K\"unneth decomposition $\{p_i : 0 \leq i \leq 8\}$ of
$X$ given in Theorem \ref{theoremCK} which, by Proposition
\ref{specialMurre}, is a Murre decomposition. Let $C$ be a smooth
projective curve and let $\{p_0^C, p_1^C, p_2^C\}$ be a self-dual
Chow--K\"unneth decomposition for $C$ as described in \cite{Scholl}.
Then the variety $X \times C$ has a self-dual Chow--K\"unneth
decomposition given by $q_l := \sum_{i+j=l} p_i \times p_j^C$. The
results of \cite{Vial2} make it possible to prove the following.
\begin{theorem}
The fivefold $X \times C$ endowed with the above self-dual
Chow--K\"unneth decomposition satisfies Murre's conjectures (A), (B)
and (D).
\end{theorem}
\begin{proof} The idempotents $p_0$, $p_1$, $p_2$, $p_3$, $p_0^C$ and
$p_1^C$ factor through varieties of respective dimension $0$,
$1,2,1,0$ and $1$. Therefore, $(X,q_i)$ is isomorphic to a direct
summand of the twisted motive of a surface for $i$ even. It is
isomorphic to the direct summand of the twisted motive of a curve
for $i = 1$ or $9$, and it is isomorphic to the direct summand of
the twisted motive of a threefold for $i$ odd $\neq 1, 9$. Murre's
conjectures (B) and (D) for $X$ endowed with the Chow--K\"unneth
decomposition given by the $q_l$'s then follow immediately from
Proposition \ref{MurreB} and from the first point of Proposition
\ref{MurreC}.
\end{proof}
\section{Application to the finite-dimensionality problem}
\label{fdprob}
Kimura \cite{Kimura} introduced the notion of finite-dimensionality
for Chow motives. There he proved that any variety dominated by a
product of curves is finite-dimensional. It was proved by Guletskii
and Pedrini \cite{GP} that a surface with representable Chow group of
zero-cycles is Kimura finite-dimensional. Gorchinskiy and Guletskii
\cite{GG} then proved that a threefold with representable Chow group
of zero-cycle is Kimura finite-dimensional. This was subsequently
generalised to varieties of any dimension in \cite{Vial3} and to pure
motives in \cite{Vial1}. In their paper, Gorchinskiy and Guletskii
also prove \cite[Theorem 15]{GG} that, when $X$ is fibred over a curve
by del Pezzo or Enriques surfaces over an algebraically closed field
of characteristic zero, then $X$ has representable Chow group of
zero-cycles. Their method involves looking at the singular fibres of
the family. Our Theorem \ref{isogeny} is more general and immediately
gives
\begin{theorem} \label{surface-Kimura} Let $X$ be a smooth projective
threefold over a field $k$ and let $f : X \r C$ be a dominant
morphism over a curve $C$ such that the general fibre of $f_\Omega$
has trivial Chow group of zero-cycles. Then $X$ has representable
Chow group of zero-cycles and is finite-dimensional in the sense of
Kimura.
\end{theorem}
Godeaux surfaces are examples of surfaces of general type with trivial
Chow group of zero-cycles \cite{Godeaux}. Therefore, new cases
encompassed by the above theorem are given by threefolds fibred by
Godeaux surfaces over a curve. Let's make this more precise.
Let $\zeta$ be a primitive fifth root of unity. The group $G = \Z/5\Z$
acts on the complex projective space $\P_\C^3$ in the following way :
$\zeta \cdot [x_0:x_1:x_2:x_3] = [x_0:\zeta x_1: \zeta^2 x_2:
\zeta^3x_3]$. Let $\bar{V} := H^0(\P^3,O_{\P^3}(5))^G$ be the subspace
of $H^0(\P^3,O_{\P^3}(5))$ consisting of elements invariant under the
action of $G$ and let $V \hookrightarrow \bar{V}$ be the Zariski open
subset of $\bar{V}$ consisting of elements defining smooth quintic
surfaces. The monomials $X_i^5$ belong to $\bar{V}$ so that the
dimension of $\bar{V}$ is at least $4$. If $Y_v$ is a smooth quintic
in $\P^3$ given by the equation $v \in V$, then a local computation
shows that $Y_v$ cannot contain the fixed points of the action of $G$
on $\P^3$, so that the action of $G$ restricts to a free action on
$Y_v$. The quotient space $X_v := Y_v/G$ is a smooth projective
surface called a Godeaux surface. These Godeaux surfaces fit into a
family $X \r \P (\bar{V})$.
Let's consider a smooth projective curve $C$ in $\P(\bar{V})$ that
meets $\P(\bar{V} - V)$ transversely. Then ${X}$ restricted to $C$
gives a smooth projective threefold ${X}|_C \r C$ with general fibre a
Godeaux surface. If $C$ is of general type ($g(C) \geq 2$), then, by a
result of Viehweg \cite{Viehweg} which is a special instance of the
Iitaka conjecture, ${X}|_C$ is a threefold of general type. We
have thus exhibited new examples of threefolds of general type with
representable Chow group of zero-cycles (obvious examples are given by
the product of a curve of general type with a Godeaux surface). Such
threefolds are also Kimura finite-dimensional thanks to \cite{GG}.
\medskip
In the following theorem, by conic fibration, we mean a dominant
morphism $X \r S$ whose general fibre is a conic.
\begin{theorem} \label{conic-Kimura} Let $X$ be a smooth projective
threefold which is a conic fibration over a surface $S$ which is
Kimura finite-dimensional. Then $X$ is finite-dimensional in the
sense of Kimura.
\end{theorem}
\begin{proof}
In section \ref{CK}, we proved that there is an orthogonal
decomposition of the diagonal $\Delta_X = p_0 +p_1+p_2^{tr} +
p_2^{alg} + p_3 + {}^tp_2^{alg}+ {}^tp_2^{tr}+{}^tp_1+{}^tp_0$ with
$(X,p_0)$ and $(X,p_2^{alg})$ isomorphic to twisted motives of
points, $(X,p_1)$ and $(X,p_3)$ isomorphic to direct summands of
twisted motives of curves; and with $(X,p_2^{tr})$ isomorphic to
$(S,\pi_2^{tr,S})$. Motives of points and motives of curves are
finite-dimensional \cite{Kimura}. Since $S$ is Kimura
finite-dimensional by assumption and since finite-dimensionality is
stable under direct summand \cite{Kimura}, we have that
$(X,p_2^{tr})$ is finite-dimensional. Therefore $X$ is Kimura
finite-dimensional.
\end{proof}
\begin{theorem} \label{C-Murre2}
Let $X$ be as in Theorem \ref{surface-Kimura} or as in Theorem
\ref{conic-Kimura}. Then $X$ satisfies Murre's conjectures (A),
(B), (C) and (D).
\end{theorem}
\begin{proof}
By Theorem \ref{Murre-theorem}, $X$ is Kimura finite-dimensional and
has a Chow--K\"unneth decomposition that satisfies Murre's
conjectures (B) and (D). According to \cite[Theorem 4.8]{Vial2}, we
can conclude that $X$ satisfies Murre's conjecture (C) if the
cohomology of $X$ is generated via the action of correspondences by
the cohomology of surfaces. By Theorem \ref{Murre-theorem} again,
$X$ satisfies the Lefschetz standard conjecture. Therefore, it
suffices to show that $H_3(X)$ is generated by the $H_1$ of a curve.
But then, this follows, via Bloch--Srinivas \cite{BS}, from the fact
that $\CH_0(X_\Omega)$ is supported on a surface.
\end{proof}
\section{A fourfold of general type satisfying Murre's conjectures}
\label{general}
In this section we wish to give explicit examples of fourfolds
satisfying the assumptions of Theorem \ref{mainth}. For this purpose,
we consider two-dimensional families of surfaces having trivial Chow
group of zero-cycles.
A first type of such families was already given in Theorem
\ref{ratconn2} and consisted in separably rationally connected
fibrations over a surface (separably rationally connected is the same
as rationally connected if the base field has characteristic zero).
Precisely, Theorem \ref{ratconn2} considered smooth projective
varieties $X$ over a field $k$ with an equidimensional map $X \r S$ to
a smooth projective surface with general fibre being separably
rationally connected.
However, such a fourfold is not of general type. A natural question is
to ask whether it is possible to construct a fourfold of general type
that has a self-dual Murre decomposition. In \cite[\S 2.3 \& Cor
4.12]{Vial2}, we produced examples of such fourfolds. These fourfolds
have the property of having trivial Chow group of zero-cycles. Obvious
examples were given by the product of two surfaces of general type
with trivial Chow group of zero-cycles (e.g. Godeaux surfaces).
Another example, a fourfold of Godeaux type, was given. The strategy
consisted in checking the validity of the generalised Hodge conjecture
for this fourfold.
We are now going to give an example of fourfold of general type with
non-trivial (and in fact non-representable) Chow group of zero-cycles
that has a self-dual Murre decomposition. Let's take up the notations
of the previous section and let's consider the family $X \r
\P(\bar{V})$. Let then $S$ be a high-degree (i.e. $\geq 5$) complete
intersection which is a smooth surface in $\P(\bar{V})$ meeting
$\P(\bar{V}-V)$ transversely. Then $X|_S$ is a projective fourfold
with $X|_S \r S$ having a smooth Godeaux surface as a general fibre. A
desingularization $X' \r X|_S$ gives a morphism $X' \r S$ with general
fibre being of general type and having trivial Chow group of
zero-cycles. This is because these two conditions are birational
invariants. The high-degree condition on $S$ imposes that $S$ is of
general type and has non-representable $\CH_0(S)_\alg$. Therefore, by
Viehweg's result \cite{Viehweg}, $X'$ is of general type; and, by
Theorem \ref{mainth}, $X'$ has a self-dual Murre decomposition which
satisfies the motivic Lefschetz conjecture.
\section{Application to unramified cohomology} \label{unram}
Following the fundamental result of Colliot-Th\'el\`ene, Sansuc and
Soul\'e \cite{CTSS} which asserts that the degree-three unramified
cohomology groups $H^3_{nr}(S/k,\Q_l/\Z_l(2))$ vanish for all prime
numbers $l$ for $S$ a smooth projective surface defined over a field
$k$ which is either finite or separably closed, it is proved in
\cite[Proposition 3.2]{CTK} that, if $X$ is a smooth projective variety
defined over a field $k$ which is either finite or separably closed
such that its Chow group of zero-cycles is supported on a surface,
then the groups $H^3_{nr}(X/k,\Q_l/\Z_l(2))$ are finite for all prime
numbers $l$ and vanish for almost all $l$. Therefore, any variety $X$
defined over a finite field or a separably closed field such that its
restriction $X_\Omega$ to a universal domain $\Omega$ satisfies the
assumptions of Theorem \ref{isogeny} has finite degree-three
unramified cohomology $\bigoplus_l H^3_{nr}(X/k,\Q_l/\Z_l(2))$. In
particular, the fourfold of general type of section \ref{general}, when
defined over a finite field or a separably closed field, has finite
degree-three unramified cohomology. Furthermore, as a straightforward
application of Theorem \ref{repsurface}, we get
\begin{proposition} \label{bielliptic-unram} Let $f : X \r C$ be a
dominant and generically smooth morphism from a smooth projective
variety $X$ to a smooth projective curve $C$ defined over a field
$k$ which is either finite or separably closed. Assume that the
general fibre $Y$ of $f_\Omega$ is such that $\CH_0(Y)_\alg$
is representable. Then $H^3_{nr}(X/k,\Q_l/\Z_l(2))$ is finite for
all prime numbers $l$ and vanishes for almost all $l$.
\end{proposition}
Since unramified cohomology is a birational invariant for smooth
projective varieties, the conclusion of the above theorem still holds
for a smooth projective variety $X'$ which is birational to the
variety $X$ of the theorem. For instance, we get finiteness of degree
three unramified cohomology for threefolds which are the smooth
compactification of one-dimensional families of smooth projective
surfaces defined over a finite field or a separably closed field whose
generic member is a bielliptic surface.
\begin{footnotesize}
\bibliographystyle{plain}
|
\section{Introduction}
Strong interaction is known to respect space and time reflection symmetry to a very high degree.
However this is not a direct consequence of laws of quantum chromodynamics (QCD), which, in
principle permit a parity violating term or the so called $\theta$-term given as
\begin{equation}
{\cal L}_\theta=\frac{\theta}{64\pi^2} g^2 F_{\mu\nu}^a\tilde F^{a\mu\nu}.
\label{lth}
\end{equation}
In the above, $F_{\mu\nu}^a$ is the gluon field strength and $\tilde F^{\mu\nu}$ being its dual.
This term while being consistent with Lorentz invariance and gauge invariance, it violates charge
conjugation and parity unless $\theta=0$ mod $\pi$. However, CP symmetry conserving nature of QCD
has been established by precise experiments that sets limit on the intrinsic electric dipole moment
of neutron. The current experimental limit on this leads to limit on the coefficient of the CP
violating term of the QCD Lagrangian as $\theta < 0.7\times 10^{-11}$ \cite{endm}. This smallness
of the CP violation term or its complete absence is not understood completely though a possible
explanation is given in terms of spontaneous breaking of a new symmetry : the Peccei-Quinn symmetry
\cite{pecceiquinn} which could give rise to axions. For zero temperature and and zero density,
spontaneous parity violation does not arise for $\theta=0$ by the well known Vafa-Witten theorem
\cite{vafawit}. On the other hand for $\theta=\pi$ there could be spontaneous CP violation by the
so called Dashen phenomena \cite{dashen}. Because of the nonperturbative nature of this $\theta$
term in QCD, this problem has been studied extensively in low energy effective theories like chiral
perturbation theory \cite{chpt}, linear sigma model \cite{fragalsm} as well as Nambu-Jona-Lasinio
(NJL) model and its different extensions \cite{cpnjl,bbone,bbtwo,sakai}.
Even if CP is not violated for QCD vacuum, it is possible that it can be violated for QCD matter at
finite temperature or density. It has been proposed that hot matter produced in heavy ion collision
experiments can give rise to domains of meta stable states that violate CP \cite{dimacp}.
Experimental signatures for the existence of local CP violation has been based on charge separation
of hadronic matter due to the strong magnetic field produced in heavy ion collision experiments by
a mechanism called chiral magnetic effect (CME) \cite{cme}. This mechanism may explain the charge
separation in the recent STAR results \cite{starexp}.
In the present work we focus our attention on how chiral transition is affected when there is CP
violating term in the Lagrangian. For this purpose, we adopt the 3-flavor NJL model as an effective
theory for chiral symmetry breaking in strong interaction \cite{klevansky,rehberg}. The CP violating
parameter $\theta$ is included in the Kobayashi-Maskawa-t'Hooft (KMT) determinant term. In this
context we note that the two flavor scenario for spontaneous CP violation for $\theta=\pi$ has been
studied in this model \cite{bbone}. This has been further extended to study the restoration of CP at
finite temperature \cite{bbtwo}. The effect of the theta vacuum on the deconfinement and chiral
transition has also been analyzed within a two flavor NJL model with Polyakov loop \cite{sakai}.
We organize the present work as follows. In the next section we shall discuss the 3-flavor NJL model
with a CP violating term. We consider a variational ground state with quark-antiquark pairs that is
related to chiral symmetry breaking. The ansatz functions are to be determined through minimization
of the thermodynamic potential. The ansatz is general enough to include both scalar as well as
pseudoscaler condensates. As we shall see the pseudoscalar condensates develop for non zero values
of $\theta$ in the KMT determinant term. In section III we discuss the resulting phase diagram at
finite temperature as well as finite density for different values of the CP violating parameter in
the Lagrangian. In Section IV we summarize our results and give a possible outlook.
\section{NJL model with CP violation and an ansatz for the ground state}
To describe the chiral phase structure of strong interactions including the CP violating effects,
we use the 3-flavor NJL model along with the flavor mixing determinant term. The Lagrangian is
given by
\begin{equation}
{\cal L} = \bar\psi\left(i\partial\hspace{-0.15cm}/ - m\right)\psi
+ G\sum_{A=0}^8\left[(\bar\psi\lambda^A\psi)^2 + (\bar\psi i\gamma^5\lambda^A\psi)^2\right]
- K\left[e^{i\theta}det\lbrace\bar\psi(1+\gamma^5)\psi\rbrace +
e^{-i\theta}det\lbrace\bar\psi(1-\gamma^5)\psi\rbrace\right],
\label{lag3fl}
\end{equation}
\noindent
where $\psi ^{i,a}$ denotes a quark field with color `$a$' $(a=r,g,b)$, and flavor `$i$'
$(i=u,d,s)$, indices. The matrix of current quark masses is given by $\hat m$ =
diag$_f(m_u,m_d,m_s)$ in the flavor space. We shall assume in the present investigation, isospin
symmetry with $m_u$ = $m_d$. In Eq.(\ref{lag3fl}), $\lambda^A$, $A=1,\cdots 8$ denote the
Gell-Mann matrices acting in the flavor space and
$\lambda^0 = \sqrt{\frac{2}{3}}\,1\hspace{-1.5mm}1_f$, where $\,1\hspace{-1.5mm}1_f$ is the unit
matrix in the flavor space. The four point interaction term $\sim G$ is symmetric under
$SU(3)_V\times SU(3)_A\times U(1)_V\times U(1)_A$. The determinant term $\sim K$, which generates
a six point interaction for the case of three flavors, breaks $U(1)_A$ symmetry for vanishing
$\theta$ values. The effect of topological term of Eq.(\ref{lth}) is simulated by the determinant
term of Eq.(\ref{lag3fl}) in the quark sector.
We shall next consider an ansatz for the ground state with quark-antiquark condensates which
includes both the scalar as well as CP violating pseudoscalar channel. To make notations clear,
we first write down the field operator expansion for the quark fields as given in
\cite{hmspmnjl,amspm},
\begin{equation}
\psi (\zbf x,t=0 )\equiv \frac{1}{(2\pi)^{3/2}}\int \tilde\psi(\zbf k)
e^{i\zbf k\cdot\zbf x}d\zbf k
=\frac{1}{(2\pi)^{3/2}}\int \left[U_0(\zbf k)q^0(\zbf k )
+V_0(-\zbf k)\tilde q^0(-\zbf k )\right]e^{i\zbf k\cdot \zbf x}d \zbf k,
\label{psiexp}
\end{equation}
\noindent
where $U^0(\zbf k)$ and $V^0(-\zbf k)$ are the four component spinors which can be explicitly
written as,
\begin{eqnarray}
U_0(\zbf k )=&&\left(\begin{array}{c}\cos(\frac{\chi^0}{2})\\
\zbf \sigma \cdot \hat k \sin(\frac{\chi^0}{2})
\end{array}\right)\;\;\;\;\;\;\;\;and\;\;\;\;\;\;\;\;
V_0(-\zbf k )=
\left( \begin{array}{c} -\zbf \sigma \cdot \hat k \sin(\frac{\chi^0}{2})
\\ \cos(\frac{\chi^0}{2})\end{array}
\right).
\label{uv0}
\end{eqnarray}
The superscript `$0$' indicates that the operators $q^0$ and $\tilde q^0$ are the two component
operators for the quark annihilation and antiquark creation corresponding to the the perturbative
or the chiral vacuum $|0\rangle$. Here we have suppressed the color and flavor indices of the quark
field operators. The function $\chi^0(\zbf k)$ in the spinors in Eq.(\ref{uv0}) are given as
$\cot{\chi_i^0}=m_i/|\zbf k|$, for free massive fermion fields, $i$ being the flavor index. For
massless fields $\chi^0(|\zbf k|)=\pi/2$.
We next consider an ansatz of the ground state at zero temperature as
\begin{equation}
|\Omega\rangle=U_q|0\rangle,
\end{equation}
where, $U_q=U_{qI}U_{qII}$ is an unitary operator. $U_{qI}$ and $U_{qII}$ are unitary operators
described in terms of quark-antiquark creation and annihilation operators. Explicitly they are
given as
\begin{equation}
U_{qI}=\exp\left(\int d\zbf k q_r^0(\zbf k)^\dagger (\bfm\sigma\cdot\hat{\zbf k})_{rs}
f(k)\tilde q_{s}^0(-\zbf k)-h.c\right)
\label{uqI}
\end{equation}
and
\begin{equation}
U_{qII}=\exp\left(\int d\zbf k q_r(\zbf k)^\dagger
r g(k)\tilde q_{-r}(-\zbf k)-h.c\right)
\end{equation}
where $f(k)$ and $g(k)$ are the ansatz functions which are to be determined later from the
extremization of the thermodynamic potential.
Finally, to include the effect of temperature and baryon density, we use the techniques of
thermofield dynamics (TFD) which is quite convenient while dealing with operators and states
\cite{tfd,amph4}. Here, the statistical average of an operator is given as an expectation value
over a `thermal vacuum'. The methodology of TFD involves the doubling of the Hilbert space
\cite{tfd}. Explicitly, the `thermal vacuum' is constructed from the ground state at zero
temperature and density through a thermal Bogoliubov transformation given as
\begin{equation}
|\Omega(\beta,\mu)\rangle = {\cal U}_F|\Omega\rangle =
e^{{\cal B}(\beta,\mu)^\dagger-{\cal B}(\beta,\mu)}|\Omega\rangle
\label{ubt}
\end{equation}
with,
\begin{equation}
{\cal B}^\dagger(\beta,\mu)=\int \Big [ \theta_-(\zbf k, \beta,\mu)
q^\prime (\zbf k)^\dagger \underline q^{\prime} (-\zbf k)^\dagger +
\theta_+(\zbf k, \beta,\mu) \tilde q^\prime (\zbf k)
\underline { \tilde q}^{\prime} (-\zbf k)
\Big ] d\zbf k.
\label{bth}
\end{equation}
In Eq.(\ref{bth}) the ansatz functions $\theta_{\pm}(\zbf k,\beta,\mu)$ will be related to the
quark and the antiquark thermal distributions respectively and the underlined operators are the
operators in the extended Hilbert space associated with thermal doubling in TFD method.
In the following section we shall compute the thermodynamic potential which will involve
calculating thermal average of different operators which is given by the expectation values of the
corresponding operator with respect to the state given in Eq.(\ref{ubt}). This can be evaluated
directly by realizing that the state $|\Omega(\beta,\mu)\rangle$ is obtained from the state
$|0\rangle$ by successive Bogoliubov transformation. Thus e.g. we have
\begin{equation}
\langle \Omega(\beta\mu)|
\psi^{\dagger i a}_\alpha (\zbf x)\psi^{j b}_\beta (\zbf y)
|\rangle\Omega(\beta\mu)\rangle=\delta^{ij}\delta^{ab}\int\frac{d\zbf k}{(2\pi)^3}
e^{-i\zbf k\cdot(\zbf x-\zbf y)}\Lambda^i(\zbf k,\beta,\mu)_{\beta\alpha}
\label{master}
\end{equation}
where,
\begin{equation}
\Lambda^i(\zbf k,\beta,\mu) = \frac{1}{2}\left[(\cos^2\theta^i_++\sin^2\theta_-^i)
+ (\sin^2\theta^i_--\sin^2\theta_+^i)(\gamma^0\cos\phi^i\cos 2g^i+\zbf\alpha\cdot\hat
{\zbf k}
\sin\phi^i\cos 2g^i-i\gamma^0\gamma^5\sin 2g^i\right]
\end{equation}
where, we have introduced a new function $\phi^i(\zbf k)=\chi_0+2 f^i(\zbf k)$ in terms of the
condensate function $f^i(\zbf k)$ of Eq.(\ref{uqI}). From Eq.(\ref{master}), it is easy to
calculate the scalar and pseudoscalar condensates. In terms of the ansatz functions
$\phi^i(\zbf k)$ and $g^i(\zbf k)$ the scalar and pseudoscalar condensates for the i-th flavor
can be respectively written as
\begin{eqnarray}
\langle\bar\psi\psi\rangle_i &=&
-\frac{2N_c}{(2\pi)^3}\int d\zbf k\cos \phi^i\cos{2g^i}(1-n^i_- - n^i_+) \equiv -I_s^i\\
\label{scalar}
\langle\bar\psi\gamma_5\psi\rangle_i &=&
-i\frac{2N_c}{(2\pi)^3}\int d\zbf k\sin{2g^i}(1-n^i_- - n^i_+) \equiv
-iI_p^i
\label{pseudo}
\end{eqnarray}
where $n^i_\mp=\sin^2\theta^i_\mp$. Thus a non vanishing $I_s^i$ will imply chiral symmetry
breaking phase while a non vanishing $I_p^i$ or equivalently $g^i(\zbf k)$ will indicate CP
violating phase. The condensate functions $\phi^i(\zbf k)$, $g^i(\zbf k)$ as well as the thermal
functions $\theta^i_\mp(\zbf k,\beta,\mu)$ shall be determined by extremization of the
thermodynamic potential with respect to the respective functions. We shall carry out these
extremization in the following section.
\section{Evaluation of thermodynamic potential and gap equations }
\label{evaluation}
As mentioned we shall be considering the chiral phase structure in presence of the CP violating
terms within the frame work of Nambu-Jona-Lasinio model of Eq.(\ref{lag3fl}). The energy density
is given by the expectation value of the Hamiltonian corresponding to the Lagrangian given in
Eq.(\ref{lag3fl}) with respect to the thermal ansatz of Eq.(\ref{ubt}). The energy density can be
written as
\begin{equation}
\epsilon = T + V = T + V_S + V_D
\label{entot}
\end{equation}
where $T$ is the expectation value of the kinetic term in Eq.(\ref{lag3fl}) and using
Eq.(\ref{master}), is given as
\begin{equation}
T = \langle\psi^\dag(-i\vec\alpha\cdot\vec\nabla+\beta m)\psi\rangle
= -\frac{2N_c}{(2\pi)^3}\sum_i\int{d\zbf k(m^i\cos \phi^i + |\zbf k|\sin \phi^i)
\cos{2g^i}(1-n^i_--n^i_+)}.
\label{kinetic}
\end{equation}
$V_S$ the contribution from the four point interaction term in Eq.(\ref{lag3fl}) to the energy
density and using Eq.(\ref{master}), this is given as
\begin{equation}
V_S = -G\langle\sum_{A=0}^8\left[(\bar\psi\lambda^A\psi)^2 +
(\bar\psi i\gamma_5\lambda^A\psi)^2\right]\rangle
= -2G\sum_i\left[{I_s^i}^2 + {I_p^i}^2\right].
\label{vs}
\end{equation}
Finally, $V_D$ denotes the contribution from the determinant interaction term in Eq.(\ref{lag3fl}) and
using Eq.(\ref{master}), is given as
\begin{eqnarray}
V_D &=& K\langle e^{i\theta}det\lbrace\bar\psi(1+\gamma_5)\psi\rbrace +
e^{-i\theta}det\lbrace\bar\psi(1-\gamma_5)\psi\rbrace\rangle \nonumber\\
&=& 2K\left[\cos\theta\left\lbrace-\prod_{i=1}^3I_s^i +
\frac{1}{2}|\epsilon_{ijk}|I_s^i I_p^j I_p^k\right\rbrace
+ \sin\theta\left\lbrace-\prod_{i=1}^3I_p^i +
\frac{1}{2}|\epsilon_{ijk}|I_s^i I_s^j I_p^k\right\rbrace\right].
\label{vd}
\end{eqnarray}
Now, the thermodynamic potential is given as
\begin{equation}
\Omega = \epsilon -\mu \rho- \frac{S}{\beta},
\label{omega1}
\end{equation}
where $\mu$ is the quark chemical potential corresponding to the quark number density $\rho$
given as
\begin{equation}
\rho = \sum_{i=u,d,s}\langle\psi^\dag\psi\rangle_i =
\frac{2N_c}{(2\pi)^3}\sum_{i=u,d,s}\int{d\zbf k(1-\sin^2\theta^i_+ + \sin^2\theta^i_-)}.
\label{numden}
\end{equation}
Finally, $S$ is the entropy density given as
\begin{equation}
S =\frac{2N_c}{(2\pi)^3}\sum_{i=u,d,s}
\int d\zbf k\left(\cos^2\theta_-^i\ln \cos^2\theta_-^i+
\sin^2\theta_-^i\ln\sin^2\theta_-^i
+\cos^2\theta_+^i\ln \cos^2\theta_+^i
+ \sin^2\theta_+^i\ln\sin^2\theta_+
\right).
\label{sq}
\end{equation}
Thus the thermodynamic potential given in Eq.(\ref{omega1}) is known in terms of the ansatz
functions of Eq.(\ref{ubt}). Extremizing the thermodynamic potential with respect to
$\phi^i(\zbf k)$ and $g^i(\zbf k)$ respectively leads to
\begin{eqnarray}
\tan \phi^i = \frac{|\zbf k|}{M_s^i}\;\;\;\;\;\;\;\;and\;\;\;\;\;\;\;\;
\tan{2g^i} = \frac{M_p^i}{\sqrt{{M_p^i}^2 + |\zbf k|^2}}.
\end{eqnarray}
where $M_s^i$ and $M_p^i$ are respectively the contributions to the constituent quark mass
(for i-th flavor) from the scalar and pseudoscalar condensates and they are given by,
\begin{eqnarray}
M_s^i &=& m^i + 4GI_s^i + K|\epsilon_{ijk}|\lbrace\cos\theta(I_s^j I_s^k - I_p^j I_p^k) -
\sin\theta(I_s^j I_p^k + I_p^j I_s^k)\rbrace \\
\label{scmass}
M_p^i &=& 4G I_p^i - K|\epsilon_{ijk}|\lbrace\cos\theta(I_s^j I_p^k + I_p^j I_s^k) -
\sin\theta(I_p^j I_p^k - I_s^j I_s^k)\rbrace.
\label{psmass}
\end{eqnarray}
Finally extremizing the thermodynamic potential with respect to the thermal
distribution functions $\theta^i_{\mp}$ leads to
\begin{equation}
\sin^2\theta^i_\pm=\frac{1}{\exp(\omega^i\mp\mu^i)+1},
\label{sth}
\end{equation}
where, $\omega^i(\zbf k)=\sqrt{\zbf k^2+({M_s^i}^2+{M_p^i}^2)}$. Thus we can see that the
constituent quark masses get contribution from both the scalar and pseudoscalar condensates.
Substituting the extremized solution for the condensate functions $\tan\phi^i$ and $\tan{2g^i}$
in Eq.(\ref{scalar}) and Eq.(\ref{pseudo}), we have the self consistent equations for the scalar
and pseudoscalar condensates,
\begin{eqnarray}
I_s^i &\equiv& -\langle\bar\psi\psi\rangle_i =
\frac{2N_c}{(2\pi)^3}\int{d\zbf k\left(1-n_-^i-n_+^i\right)\frac{M_s^i}{\omega^i}}
\label{isi}\\
I_p^i &\equiv& i\langle\bar\psi\gamma_5\psi\rangle_i =
\frac{2N_c}{(2\pi)^3}\int{d\zbf k\left(1-n_-^i-n_+^i\right)\frac{M_p^i}{\omega^i}}.
\label{ipi}
\end{eqnarray}
Thus with the scalar and the pseudoscalar condensates given as above,
Eq.s(\ref{scmass}-\ref{psmass}) are actually coupled self-consistent equations for $M_s^i$ and
$M_p^i$.
Substituting the extremized solutions for the condensate functions and using the gap equations
Eq.(\ref{scmass}-\ref{psmass}) in Eq.(\ref{omega1}), the thermodynamic potential becomes
\begin{eqnarray}
\Omega &=& -\frac{2N_c}{(2\pi)^3}\sum_i\int{d\zbf k(\omega^i-|\zbf k|)} +
2G_s\sum_i\left[{I_s^i}^2 - {I_p^i}^2\right] + \sum_i M_p^i I_p^i\nonumber\\
&&+ 4K\cos\theta\prod_{i=1}^3I_s^i - 2K\sin\theta\left[\prod_{i=1}^3I_p^1 +
\frac{1}{2}|\epsilon_{ijk}|I_s^i I_s^j I_p^k\right] \nonumber \\
&&- \frac{2N_c}{\beta(2\pi)^3}\sum_i\int{d\vec k\left[\ln{\lbrace 1+e^{-\beta(\omega^i-\mu^i)}\rbrace} +
\ln{\lbrace 1+e^{-\beta(\omega^i+\mu^i)}\rbrace}\right]}.
\label{thermpot}
\end{eqnarray}
In the above we have subtracted the perturbative vacuum energy density contribution. Let us note
that, the effective potential has been calculated using an explicit ansatz for the condensate and
not evaluating the effective potential at a mean field level after performing a chiral
transformation for the quarks so as to remove it from the determinant term as it has been computed
in Ref.\cite{bbone,sakai}.
Eq.(\ref{thermpot}) for the thermodynamic potential and the gap equations for the scalar and
pseudoscalar masses i.e. Eq.(\ref{scmass}-\ref{psmass}) shall be the focus of our numerical analysis
which we do in the following section.
\section{results and discussions}
For numerical calculations, we have taken the values of the parameters of the NJL model as follows.
The coupling constant $G_s$ has the dimension of $[{\rm Mass}]^{-2}$ while the six fermion coupling
$K$ has a dimension $[{\rm Mass}]^{-5}$. To regularize the divergent integrals we use a sharp
cut-off, $\Lambda$ in 3-momentum space. Thus we have five parameters in total, namely the current
quark masses for the non strange and strange quarks, $m_q$ and $m_s$, the two couplings $G_s$, $K$
and the three-momentum cutoff $\Lambda$. We have chosen here $\Lambda=0.6023$ GeV,
$G_s\Lambda^2=1.835$, $K\Lambda^5=12.36$, $m^q=5.5$ MeV and $m^s=0.1407$ GeV as has been used in
Ref.\cite{rehberg}. After choosing $m^q=5.5$ MeV, the remaining four parameters are fixed by fitting
to the pion decay constant and the masses of pion, kaon and $\eta'$. With this set of parameters the
mass of $\eta$ is underestimated by about six percent and the constituent masses of the light quarks
turn out to be $M^{u,d}=0.368$ GeV for u-d quarks, while the same for strange quark turns out as
$M^s=0.549$ GeV, at zero temperature and zero density.
For a given temperature and the chemical potential, we first solve the coupled self consistent gap
equations Eq.s(\ref{scmass}-\ref{psmass}) with the parameters of the model as above. Since we have
assumed isospin symmetry and have $m^u=m^d$, these are actually four coupled equations: two for the
scalar condensates related to the two masses $M_s^u=M_s^d$, $M_s^s$ and two equations for the
pseudoscalar condensate related to the corresponding mass parameters $M_p^u=M_p^d$, $M_p^s$. The
solutions to these equations are then substituted in Eq.(\ref{thermpot}) and checked regarding
minimum of the thermodynamic potential. If there are more solutions to the gap equation, the one
with the minimum thermodynamic potential is chosen.
Let us first discuss the ground state structure at zero temperature and zero density. In
Fig.(1-a) we show the theta dependence of contributions to the mass of up quark from the scalar as
well as the pseudoscalar condensates. As is clearly seen as $\theta$ increases the condensates in
the two channels behave in a complimentary manner. While the magnitude of scalar condensates
decreases with $\theta$ (till $\theta=\pi$), the magnitude of the pseudoscalar condensate increases
so that the total constituent quark mass $M=\sqrt{{M_s}^2+{M_p}^2}$ remains almost the same.
Spontaneous CP violation is clearly seen for $\theta=\pi$ with two degenerate solutions for $M_p^u$
differing by a sign. In Fig.(1-b) we show the effective potential as calculated above as a function
of $\theta$. The effective potential is normalized with respect to the same at $\theta=0$. The
minimum of the potential is at $\theta=0$ which is consistent with the Vafa-Witten theorem and has
a cusp at $\theta=\pi$ which has also been observed in 2-flavor NJL model \cite{bbone}.
\begin{figure}[t]
\vspace{-0.4cm}
\begin{center}
\begin{tabular}{c c}
\includegraphics[width=9cm,height=9cm]{u_mu0t0_gap_th.eps}&
\includegraphics[width=9cm,height=9cm]{mu0t0_veff_th.eps}\\
Fig. 1-a & Fig. 1-b
\end{tabular}
\end{center}
\caption{ $\theta$-dependence of the condensates (Fig.1-a) and the effective potential
at zero temperature and zero baryon density (Fig.1-b).}
\label{fig1}
\end{figure}
Next we consider the effect of nonzero density and temperature. In Fig.(\ref{fig2}), we show the
variation of masses of the quarks with chemical potential at zero temperature. In Fig.(2-a) we
show the variation of masses for the case $\theta=0$. In this case the pseudoscalar condensates
vanish and the contribution to the masses of the quarks are from the scalar condensates only. The
(approximate) first order chiral transition takes place at $\mu\sim361$ MeV for u and d quarks
with their masses decreasing discontinuously to about $M^{u,d}\sim52$ MeV from their vacuum value
of $M^{u,d}=368$ MeV. Because of the flavor mixing KMT term, this decrease is reflected also in the
decrease of the strange quark mass to $M^s=464$ MeV from its vacuum value of $M^s=549$ MeV. This
result is similar to the results obtained in the context of color superconductivity in NJL model
with a determinant term \cite{hmcsc} and in the context of chiral symmetry breaking in a similar
model \cite{bccsb}. Similarly in Fig.(2-b) we show the variation of masses of up and strange quark
as well as the variation of the contributions from scalar and pseudoscalar condensates to the
constituent mass of up quark for $\theta=\pi/2$. Here the critical chemical potential for chiral
transition is $\mu_c\sim 375$ MeV where the mass contributions from scalar and pseudoscalar
condensate become 24 MeV and 12 MeV respectively from their vacuum values of 266 MeV and 248 MeV.
The total mass for the u and d quarks become 27 MeV from its vacuum value of 364 MeV. On the
other hand the contribution to the strange mass from the pseudoscalar condensate is negligible
($\sim 12$ MeV) compared to the contribution from the scalar condensate ($\sim 548$ MeV). Because
of flavor mixing again, strange quark mass also decrease to 463 MeV at $\mu_c=375$ MeV. For
$\theta=\pi$ the scalar condensate almost vanishes but for the nonzero current quark masses while
the contribution to the constituent quark mass arises from the pseudoscalar condensate as shown in
Fig.(2-c). As the quark chemical potential is increased there is a first order transition at
$\mu_c\sim 368$ MeV. At $\mu_c$ the pseudoscalar condensate vanishes and the contribution to quark
mass arises solely from the scalar condensate which is non vanishing because of the nonzero current
quark masses.
\begin{figure}[t]
\vspace{-0.4cm}
\begin{center}
\begin{tabular}{c c c}
\includegraphics[width=6cm,height=6cm]{massth0mu.eps}&
\includegraphics[width=6cm,height=6cm]{m_t0thmpi2mu.eps}&
\includegraphics[width=6cm,height=6cm]{m_t0thmpimu.eps}\\
Fig. 2-a & Fig. 2-b & Fig. 2-c
\end{tabular}
\end{center}
\caption{Quark masses as a function of quark chemical potential at zero temperature
for $\theta=0$ (Fig.2-a), $\theta=\pi/2$ (Fig.2-b) and $\theta=\pi$ (Fig.2-c). The pseudoscalar
contribution for strange quarks is zero in this range of chemical potential.}
\label{fig2}
\end{figure}
Next we discuss the condensate variations with temperature. For $\theta=0$, at zero baryon density
the chiral crossover transition takes place for temperature about 200 MeV as may be seen in
Fig.(3-a). As $\theta$ increases, the pseudoscalar condensates starts becoming nonzero and increase
with $\theta$. For $\theta=\pi/2$, masses arising from both type of condensates are shown in
Fig.(3-b). Here, both the scalar and pseudoscalar masses show a crossover transition as temperature
is increased. In Fig.(3-c), we show the behavior of the masses for $\theta=\pi$. The transition
for the pseudoscalar mass becomes a second order transition at $\theta=\pi$ instead of a crossover
which was the nature of transition for lower $\theta$. This feature is elaborated in
Fig.(\ref{fig4}) where $\theta$ dependence of the nature of transition of scalar and pseudoscalar
condensates with temperature is shown for zero baryon density. Fig.(4-a) shows the $\theta$
dependence of the transitions for the scalar condensate. The transition is always a crossover for
scalar condensate. Fig.(4-b) shows the transitions of the pseudoscalar condensates for different
$\theta$. We can see that the transition is a second order transition for $\theta=\pi$ whereas it
is a crossover for other values of $\theta$. Similar kind of results have been obtained in sigma
model calculations \cite{fragalsm} but there the transition at $\theta=\pi$ is a first order
transition instead of a second order transition. The CP restoring transition temperature for zero
baryon density case turns out to be 192 MeV. However, the total constituent mass is nonzero as the
scalar condensate is non vanishing again because of nonzero current quark masses. This high
temperature restoration of CP is expected as the instanton effects responsible for CP violating
phase become suppressed exponentially at high temperature \cite{grosspisarski}.
\begin{figure}[htbp]
\vspace{-0.4cm}
\begin{center}
\begin{tabular}{c c c}
\includegraphics[width=6cm,height=6cm]{u_th0mu0_gap_temp.eps}&
\includegraphics[width=6cm,height=6cm]{u_thmpi2mu0_gap_temp.eps}&
\includegraphics[width=6cm,height=6cm]{u_thmpimu0_gap_temp.eps}\\
Fig. 3-a & Fig. 3-b & Fig. 3-c
\end{tabular}
\end{center}
\caption{Quark masses as a function of temperature at zero quark chemical potential
for $\theta=0$ (Fig.a), $\theta=\pi/2$(Fig b) and $\theta=\pi$.}
\label{fig3}
\end{figure}
\begin{figure}[h]
\vspace{-0.4cm}
\begin{center}
\begin{tabular}{c c}
\includegraphics[width=9cm,height=9cm]{usm_mu0temp.eps}&
\includegraphics[width=9cm,height=9cm]{upm_mu0temp.eps}\\
Fig. 4-a & Fig. 4-b
\end{tabular}
\end{center}
\caption{The nature of transition of the scalar (Fig.a) and pseudoscalar (Fig.b) condensates with
temperature at zero baryon density for different values of $\theta$.}
\label{fig4}
\end{figure}
The CP restoring transition in the present 3-flavor NJL model turns to be second order for zero
chemical potential case similar to the results for the 2-flavor case \cite{bbtwo} unlike the case
of linear sigma model coupled to quarks \cite{fragalsm}. The reason behind such different
behavior regarding the order of the transition is due to the non analytic vacuum term in the NJL
model \cite{bbtwo}. However, a first order CP transition is observed with finite chemical potential
and small temperature. This is clearly shown in Fig.(\ref{fig5}) where we have shown the dynamical
mass, $M_p^u$ arising from the pseudoscalar condensate, for different temperatures as a function
of the quark chemical potential for $\theta=\pi$. While at zero temperature, the order parameter
decreases discontinuously, as the temperature increases, it becomes less sharp and finally results
in a second order transition at high temperature.
\begin{figure}[htbp]
\vspace{-0.4cm}
\includegraphics[width=8cm,height=8cm]{upm_thmpimu.eps}
\caption{Contribution from the pseudoscalar condensates to the u-quark mass
as a function of quark chemical potential for different temperatures.
Here we have taken $\theta=\pi$.}
\label{fig5}
\end{figure}
In Fig.(\ref{fig6}), we show the phase diagram in the plane of quark chemical potential and
temperature for the CP violating transition. Since the transition is first order at zero
temperature and second order at zero chemical potential, there is a tricritical point for this
transition in this plane. This turns out to be ($\mu_c,T_c$)=(273,94) MeV. Including Polyakov loop
for the two flavor NJL model, such a tricrtical point occurs at (209,165) MeV \cite{sakai}. First
order transitions are associated with existence of meta-stable states. CP is restored in these
meta stable states and these are the nontrivial solutions of the gap equation Eq.(\ref{psmass}),
however with lower pressure than the stable solutions. In the phase diagram of Fig.(\ref{fig6}),
such solution exist in the region between solid line and the dotted line.
\begin{figure}[htbp]
\vspace{-0.4cm}
\includegraphics[width=8cm,height=8cm]{phase_diag.eps}
\caption{The phase diagram for CP transition in the $T-\mu$ plane.
The region between the solid line and the dotted line have
solutions to the pseudoscalar mass gap equation but with higher
thermodynamic potential. Here we have taken $\theta=\pi$.}
\label{fig6}
\end{figure}
\section{summary}
To summarize, we have tried to examine the effect of $\theta$ vacuum on the phase diagram of
strong interaction. This is investigated within the framework of a 3-flavor NJL model. The effect
of CP violating $\theta$ term of QCD is incorporated through the KMT determinant interaction term
in the quark space.
In the two flavor NJL model, it has been observed that spontaneous CP violation for $\theta=\pi$
occurs depending on the magnitude of current quark masses as well as the strength of the
determinant coupling \cite{bbone}. In the present case of 3-flavor NJL model, we observe that
spontaneous CP violation occurs for $\theta=\pi$ for the phenomenologically consistent parameters
\cite{rehberg} for the current quark masses as well as the strength of the determinant coupling.
To calculate the thermodynamical potential in presence of a $\theta$ term, instead of performing a
chiral rotation of the quark field operators \cite{bbone,sakai}, we have used a variational
approach using an explicit construct for the ground state. We have considered an ansatz state
general enough to have condensates both in scalar as well as pseudoscalar channel. The ansatz
functions are determined through minimization of the thermodynamic potential.
Apart from the temperature that has been considered earlier for two flavor cases
\cite{bbtwo,fragalsm} we have also considered the effect of finite quark chemical potential and
discussed the phase diagram in the $T-\mu$ plane. It turns out that for the range of temperature
and the chemical potentials we have considered, the strange quark condensates do not get
dynamically generated in the CP violating pseudoscalar channel even for non vanishing $\theta$.
None the less the strange quark antiquark condensates in the scalar channel do affect the
pseudoscalar light quark condensates through the flavor mixing coupled gap equations.
The CP restoring transition turns out to be a crossover for non vanishing values of $\theta$ that
become a second order transition for $\theta=\pi$ for zero chemical potential. However, such a
transition is a first order transition at small enough temperature and as the quark chemical
potential increases. This leads to a tricritical point in the the phase diagram for CP transition.
\def\endm{C. Baker et al., {\PRL{97}{131801}{2006}};
J. Kim and G. Carosi, Rev. Mod. Phys. {\bf 82}, 557 (2010).}
\defR. D. Peccei and H. R. Quinn, {\PRL{38}{1440}{1977}}; {\PRD{16}{1791}{1977}}.{R. D. Peccei and H. R. Quinn, {\PRL{38}{1440}{1977}}; {\PRD{16}{1791}{1977}}.}
\defC. Vafa and E. Witten, {\PRL{53}{535}{1984}}.{C. Vafa and E. Witten, {\PRL{53}{535}{1984}}.}
\defR. Dashen, {\PRD{3}{1879}{1971}}{R. Dashen, {\PRD{3}{1879}{1971}}}
\def\chpt{P. Vecchia and G. Veneziano, {\NPB{171}{253}{1980}};
A. Smilga, {\PRD{59}{114021}{1999}}; M. Tytgat, {\PRD{61}{114009}{2000}};
G. Akemann, J. Lenaghan and K. Splittorff, {\PRD{65}{085015}{2002}};
M. Creutz, {\PRL{92}{201601}{2004}};
M. Metlitski and A. Zhitnitsky, {\NPB{731}{309}{2005}}; {\PLB{633}{721}{2006}}.}
\defA. Mizher and E. Fraga, {\NPA{820}{247c}{2009}}; {\NPA{831}{91}{2009}}.{A. Mizher and E. Fraga, {\NPA{820}{247c}{2009}}; {\NPA{831}{91}{2009}}.}
\defT. Fujihara, T. Inagaki and D. Kimura, {\PTP{117}{139}{2007}}.{T. Fujihara, T. Inagaki and D. Kimura, {\PTP{117}{139}{2007}}.}
\defD. Boer and J. Boomsma, {\PRD{78}{054027}{2008}}.{D. Boer and J. Boomsma, {\PRD{78}{054027}{2008}}.}
\defD. Boer and J. Boomsma, {\PRD{80}{034019}{2009}}.{D. Boer and J. Boomsma, {\PRD{80}{034019}{2009}}.}
\defY. Sakai, H. Kouno, T. Sasaki and M. Yahiro, {\PLB{705}{349}{2011}}.{Y. Sakai, H. Kouno, T. Sasaki and M. Yahiro, {\PLB{705}{349}{2011}}.}
\defD. Kharzeev, Annals Phys. {\bf 325}, 205 (2010).{D. Kharzeev, Annals Phys. {\bf 325}, 205 (2010).}
\def\cme{D. Kharzeev, {\PLB{633}{260}{2006}};
D. Kharzeev, L. McLerran and H. Warringa, {\NPA{803}{227}{2008}};
K. Fukushima, D. Kharzeev and H. Warringa, {\PRD{78}{074003}{2008}};
K. Fukushima, M. Ruggieri and R. Gatto, {\PRD{81}{114031}{2010}}.}
\def\starexp{B. Abelev et al. [STAR Collaboration], {\PRL{103}{251601}{2009}};
{\PRC{81}{054908}{2010}}.}
\defS. Klevansky, Rev. Mod. Phys. {\bf 64}, 649 (1992).{S. Klevansky, Rev. Mod. Phys. {\bf 64}, 649 (1992).}
\defP. Rehberg, S. P. Klevansky and J. Huefner, {\PRC{53}{410}{1996}}.{P. Rehberg, S. P. Klevansky and J. Huefner, {\PRC{53}{410}{1996}}.}
\defA. Mishra and S. P. Misra, {\ZPC{58}{325}{1993}}.{A. Mishra and S. P. Misra, {\ZPC{58}{325}{1993}}.}
\defH. Mishra and S. P. Misra, {\PRD{48}{5376}{1993}}.{H. Mishra and S. P. Misra, {\PRD{48}{5376}{1993}}.}
\def\tfd{H. Umezawa, H. Matsumoto and M. Tachiki {\it Thermofield dynamics
and condensed states} (North Holland, Amsterdam, 1982);
P. A. Henning, Phys. Rep. {\bf 253}, 235 (1995).}
\defA. Mishra and H. Mishra, {\JPG{23}{143}{1997}}.{A. Mishra and H. Mishra, {\JPG{23}{143}{1997}}.}
\defA. Mishra and H. Mishra, {\PRD{74}{054024}{2006}}.{A. Mishra and H. Mishra, {\PRD{74}{054024}{2006}}.}
\defB. Chatterjee, A. Mishra, H. Mishra, {\PRD{84}{014016}{2011}}.{B. Chatterjee, A. Mishra, H. Mishra, {\PRD{84}{014016}{2011}}.}
\defD. Gross, R. Pisarski and L. Yaffe, Rev. Mod. Phys. {\bf 53}, 43 (1981).{D. Gross, R. Pisarski and L. Yaffe, Rev. Mod. Phys. {\bf 53}, 43 (1981).}
|
\section{Introduction}
The next generation of gravitational wave ({GW}) detectors - most
notably Advanced LIGO \cite{aligo}, GEO-HF \cite{geohf} and Advanced VIRGO \cite{avirgo} -
are expected to achieve the first direct detection of {GW} s
during 2015, when Advanced LIGO begins operation or soon thereafter,
giving life to the new field of the
{GW}~astronomy.
The impact on cosmology, fundamental physics, and of course on
conventional astronomy itself will be enormous; the interest in
{GW}~observation is therefore
growing rapidly in the scientific community. But while the public
interest in cosmology and relativity is high, public knowledge of the
principles behind {GW}~detection is still rather limited, and even the
existence of the large {GW}~observatories is little known by the more
general public, compared to larger experimental facilities such as,
for instance, the Large Hadron Collider \cite{LHC}. There is a clear
need to better inform and inspire the general public and prospective
students in {GW}~astronomy and related sciences.
Within the LIGO Scientific Collaboration (LSC)
the `Education and Public Outreach' (EPO) group aims to combine ideas and
approaches across the collaboration to successfully communicate the
vision and benefits of GW observation throughout the world.
To contribute to these international efforts in the promotion of GW
astronomy, at the University of Birmingham we have established a
unique program aimed to the development of small educational
computer applications that can be used to illustrate the basics of
{GW}~science and {GW}~detector technology in a playful but informative
way. The aim is to present GW science to younger generations within
one of the environments they are more familiar with, i.e. computer-games,
and to exploit the communication channels that the {\it new
technologies} offer to possibly get an even larger international
audience in contact with {GW}~science.
This activity led us to the successful development of a number of
interactive computer applets describing a variety of concepts
connected to {GW}~science and, eventually, to the realisation of
two full-scale computer-games based on the subjects of gravity and {GW}
s. In this article we overview our computer related outreach activity
and present and discuss our {GW}~related games, {Black Hole Pong}~and {Space Time Quest}.
\section{{`Processing'} programs for science outreach}
The idea of developing small computer applications for educational
purposes builds upon the need to introduce new students to the world of
computer programming and modelling of physical systems in a manner
which is enjoyable. This will not only help them learn successfully, but
also engages them with GW research.
During the initial `induction'
phase, undergraduate or summer-students involved in research projects
with our group are encouraged to generate a small computer program on
a {GW}~subject of their interest, which is then developed, as a
conventional student-project, by the student with the supervision of
more senior members of the group.
The small computer programs, called `sketches', are developed using
the open-source programming environment Processing \cite{processing:book,processing:web},
originally developed at the MIT Media Lab in 2001 as a software prototyping environment and to teach
fundamentals of computer programming within a graphical context.
Processing has eventually reached a wide audience
and is now largely used in many professional communities, such
as designers, artists, and architects to create graphical
applications, animations, interactive tools and for visual arts in
general.
Processing offers an intuitive approach to programming for
the beginner or an efficient sketchbook for rapid prototyping by
experienced programmers. Thus Processing allows students with very different
computing-backgrounds to collaborate and to successfully produce graphically
impressive sketches in a relatively short period of time.
The successful sketches are published online on our outreach website
\href{http://www.gwoptics.org}{gwoptics.org}~\cite{gwoptics}, on individual webpages where the
student-author can provide instructions and a short description of the
physics illustrated in the sketch. The program is either embedded
within the HTML code to run as an applet in the webpage itself or,
where more appropriate, distributed for download as an application to
install and run on the computer of the interested person.
The collection of sketches developed so far covers a wide spectrum of
subjects related to {GW}~science and technologies. The programs range
from illustrating the most fundamental properties of {GW} s, for
example the deformation of space-time produced by a propagating {GW}~or
the characteristic `sound' of the {GW}~signal of colliding black holes,
to the illustration of the vital technologies and phenomena used in GW
detectors such as lasers, vibration isolation systems or interference
of light. Different combinations of these sketches have been
successfully used as interactive tools during seminars about GW
detection and during more general lectures in schools and
universities, and the sketches webpages are regularly consulted online
by people interested in learning about the specific subject.
\begin{figure}[th]
\begin{center}
\includegraphics[width=7.8cm]{bhp_screen01.png}
\includegraphics[width=7.8cm]{bhp_screen03.png}
\caption{\label{fig:BHPscreenshots} Screen-shots of the {Black Hole Pong}~game. The
image on the left shows the start-up screen of the game and the
right snapshot has been
taken mid-game, showing one of the
astronomical background images, the two `black holes' controlled
by the players and several bright disks as the stars currently
in play.}
\end{center}
\end{figure}
\section{The gravitational wave games}
The positive feedback we received about the different interactive sketches has
encouraged us to realise two properly defined computer-games
based on subjects related to gravity and GW science.
The motivation behind the development of the two particular games was
as follows:
in one case, the aim was to produce an intuitive and graphically
attractive computer-game that could engage and entertain children and
teenagers during science exhibitions; in the other,
the goal was
vice-versa to develop an interactive element that could function as an
engaging and playful supplement for illustrating and explaining the
secrets of {GW}~detectors and their technology to a more educated
audience, such as high-school students onwards. The two games, {Black Hole Pong}~and
{Space Time Quest}, are presented and discussed in this section.
\subsection{Black Hole Pong}
{Black Hole Pong}~(BHP) is a new arcade-style game with a reference to
one of the very first computer-games, {\it Pong}~\cite{Pong}.
Pong involved two players, each one controlling a paddle
which they would move vertically up and down in order to bounce a
ball back towards their opponent: each
time the ball touched the opponent's far edge of the screen, the player
would score a point. In BHP the idea of the split-screen has been
kept. However each player controls a black hole which can move
horizontally as well as vertically, and the
objective is to make use of the gravitational potential of the black
hole to fling free roaming stars towards the other player.
\begin{figure}[b]
\begin{center}
\includegraphics[height=4.1cm]{BHP_commday_2011.pdf}
\includegraphics[height=4.1cm]{BHP_bsf_2010.pdf}
\includegraphics[height=4.1cm]{BHP_amaldi_2011.pdf}
\caption{\label{fig:BHPexhibition} Children, students and adults enjoying the BHP~game
at the University of Birmingham Community Day 2011 (left), the British
Science Festival 2010 (center) and the 9$^{\rm th}$ Amaldi Conference
(right).}
\end{center}
\end{figure}
BHP~has been designed as a simple, fun game for people of
all ages. At the same time, BHP features several educational elements
that make it a fascinating tool for teaching, learning and discovering
new physical concepts. For example, by learning how to manoeuvre the
incoming stars only using the black hole's gravitational potential,
the player develops an intuitive understanding of concepts such as
gravitational attraction, orbital
mechanics and gravitational slingshot effect.
The background graphics of the game is a slideshow containing images of
the night sky as well as astronomical objects taken by both amateur
astronomers and large ground/space based telescopes.
Furthermore, several astrophysical phenomena are graphically featured
in the game, such as `worm holes', `star capture'
and `gravitational lensing', adding to the overall simple but
attractive graphics, see Fig.~\ref{fig:BHPscreenshots}.
When used as part of an exhibition, science fair or similar activity we have
supported the arcade-style attraction by state-of-the art hardware:
the game is designed to be controlled with the well-known Microsoft
Xbox controllers and whenever possible we use large Apple iMac
computers to run the game, as shown in Fig.~\ref{fig:BHPexhibition}.
\subsection{Space Time Quest}
{Space Time Quest}~(STQ), here shown with screenshots in Fig.~\ref{fig:STQ:screens} and Fig.~\ref{fig:STQ:noise}, is a manager-simulation type game:
the player is the `Principal Investigator' (PI) of a
future ground-based
{GW}~interferometer who has the goal to
design
the most sensitive GW detector.
The PI is assigned a limited budget that he has to wisely distribute between
the different detector's subsystems
to tweak the instrument parameters and to achieve
the best sensitivity. Once the player is satisfied with their design,
they can operate the detector in `Science-Run' mode and see how well it performs:
the final score is determined by how far in the universe the detector itself can explore,
based on the achieved sensitivity curve,
and it is recorded in the web-based STQ `hall-of-fame'.
\begin{figure}[t]
\includegraphics[height=5.3cm]{STQ_PI.pdf}
\includegraphics[height=5.3cm]{STQ_opt.pdf}
\caption{\label{fig:STQ:screens} Screenshots of STQ. On the left, the `PI's office'
is the game's hub from which the player can access all the detector's subsystems. On the right an image of the `Optics' subsystem screen.}
\end{figure}
STQ is complemented by the \href{http://www.gwoptics.org/ebook}{`Gravitational Waves E-Book'} \cite{ebook},
effectively a collection of webpages with short introductions to a number of topics relevant to GW science.
The E-Book offers a description of the main instrument subsystems that comprise the detector and
illustrates
how each noise source relates to the
different subsystem parameters that the player chooses.
The E-Book text is purposely aimed to
the general person who has interested in GW science,
and as such it is written with a simple style
to make it accessible
to a broad public of all
ages and backgrounds.
The E-Book is also independently available online as more general reading material on GW science,
and is now also offered in multiple languages.
STQ contains many educational merits for the public.
First of all, the game showcases the physics behind a real
GW detector and it presents all the main subsystems that comprise it.
Furthermore, it illustrates all the most important noise sources
which can limit the detector's sensitivity.
By reading the E-Book and by looking at the changes in the sensitivity curve,
the player can see
how each subsystem is affected by the different noise sources,
and discover some of the ways in which physicists
try to reduce the noise sources in the detector.
The game also presents some of the typical challenges that physicists face when designing
a real physics experiment,
such
as
making trade-offs between performances of different
interlinked subsystem parameters and managing the available resources wisely.
Finally, STQ features images of astronomical objects in the background,
similar to BHP. Furthermore
the STQ graphics are based on photographs of components from real GW detectors complemented
with realistic cartoons of the detector parts, offering a realistic picture of how a GW interferometer looks like to the user.
\begin{figure}[t]
\includegraphics[height=5.3cm]{STQ_sens.pdf}
\includegraphics[height=5.3cm]{STQ_kids.pdf}
\caption{\label{fig:STQ:noise} Left: a screen-shot of the `sensitivity curve' in the STQ game.
Right: school students playing
STQ with a demonstrator during an exhibition at University of Birmingham.}
\end{figure}
STQ~is mainly targeted for science teachers, A-level, Higher and
Advanced Higher science students and is mainly suited for use in
science fairs and exhibitions and initially played with the help of
demonstrators. However, STQ~has also proved to be an entertaining
tool to teach the basics of {GW}~science also to beginners in
{GW}~research-projects and PhD schools, or to attract prospective
research students towards the {GW}~field.
\section{Use in exhibitions and distribution}
\begin{figure}[b]
\includegraphics[width=16cm]{youtube_stats.pdf}
\caption{\label{fig:youtube} Total number of unique views of the online video tutorials for the Black Hole Pong and Space Time Quest games and for one of our processing sketches, the `Augmented Reality Pendulum'. Data from \href{http://www.youtube.com/gwoptics}{youtube.com}.}
\end{figure}
Early prototypes of the games have been used for the first time
during the exhibition about {GW}~science `Looking
for Black Holes with Lasers', organised by the Birmingham GW Group
within the `British Science Festival', held in Birmingham in September
2010
\cite{BSF:main}.
The very positive feedback collected with the games
during this first exhibition gained us the attention of our university
and of local schools and associations. This led to our displays being
routinely used in University Open/Admission days and in our
university's outreach events, e.g. the University Community Day 2011
\cite{comday}, and gained members of our group invitations to visit
schools and give public seminars, where the games were used
to complement the seminar.
At the same time, positive feedback has been received from school
teachers concerning our other Processing sketches which have been used
as support material in physics lectures and during science activities.
Since their official release, BHP and STQ are freely distributed on
our website \href{http://www.gwoptics.org}{gwoptics.org}~ and on the outreach pages of the \href{http://www.ligo.org/}{ligo.org} website \cite{ligo.org}, which hosts links and multimedia material of interest for the EPO group. The launch of the two games was also announced online via
social-media networks and with online videos, with the main aim
of raising both the profile of the games and its
visibility.
Indicative figures of merit
on how positive this campaign has been
can be inferred from
the total number of
download
of the two games, the entries in the
high score `hall of fame', the unique views of the online
video-tutorials, and more generally from the number of visits to
the webpages presenting our online material.
Examples of such data are presented in Fig.\,\ref{fig:youtube} and
Fig.\,\ref{fig:gwoptics}.
The response is so far very positive and
seem to indicate a slow but constant growth of new contacts and an
increasing interest in the products themselves. In particular,
sensible increments in the number of contacts can be successfully
correlated with our contribution at science events and exhibitions,
with release of new outreach material and with our communication
campaign via online social networks.
\begin{figure}[t]
\includegraphics[width=16cm]{gwoptics_stats_cumulative_b.pdf}
\caption{\label{fig:gwoptics} Top panel: number of download of the Black Hole Pong and Space Time Quest games over the year 2011. Bottom panel, cumulative number of visits
during 2011 to some of the \href{http://www.gwoptics.org}{gwoptics.org}~outreach pages, such as the main page collecting all Processing programs, the games BHP and STQ, the STQ high-score `hall of fame', the E-Book and the pages of two other Processing sketches, the `Augmented Reality Pendulum' and the `Inspiral signal'. }
\end{figure}
STQ and BHP are also accompanied by short questionnaires, handed-out at
exhibitions and during visits to schools as well as online,
which are distributed to collect anonymous feedback amongst the users,
targeting in particular teacher's and student's categories.
The aim of the questionnaires is primarily to evaluate
the success of the games among the users and in particular about the
effectiveness of their educational aspects.
As a further step in the future,
the goal is to develop a proper analysis of the feedback provided in the questionnaires
that will allow us to better link the games to specific elements of education, such as formal education,
and to improve their integration within the syllabus.
\section{Conclusion and future activities}
Our program aimed at the development of small computer applications
for educational purposes. This has been successful, with the realisation of
several interactive sketches and two full scale games
related to GW science and technology. Thanks to our participation
at
popular science events, and to our online presence and communication
campaign, the sketches and the games are now becoming popular within
schools and science associations and, as shown by the
response gathered from the online audience, the prospective feedback looks
promising for the future. In particular,
BHP and STQ
have allowed us to significantly increase the visibility of
our online presence and as such also of
GW
science
within the general public, as well as within the local scientific community.
Encouraged by these positive results, we will continue our
computer related outreach activity in the next years, and we plan to
realise new Processing sketches for outreach in the near future.
In parallel, BHP and STQ will be treated as running projects. We will take
advantage of the feedback and advice from teachers, students and other
users to make further modifications and improvements to both games.
Furthermore, we
will continue presenting BHP, STQ and our other interactive sketches
during visits to local schools and in popular science events. We hope
to increase and improve our online communication campaign
on GW subjects to make GW detectors more and more popular among the
general public and to gain GW science the largest audience possible.
\ack
We are very grateful to the Processing community for all the online examples, code libraries and online forums
that helped us in the development of our sketches and games and without whom most of this work would not have been possible.
We thank the astronomers who allowed us to use their impressive photographs of the night sky as background images for the BHP and STQ games (see the credits page in the games for the details) and we are thankful to the GEO600 collaboration, for providing images of components of the detector used in the STQ game and for extensive beta-testing of the initial game prototype. This document has been assigned the LIGO Laboratory document number LIGO-P1100145.
\section*{References}
|
\section{Introduction}\label{sect:Intro}
Let $A$ be a Borel set in $\mathbb{R}^d$ and $\eta$ be a Poisson point process in $\mathbb{R}^d$. Assume that we observe $\eta$ and the only information about $A$ at our disposal is which points of $\eta$ lie in $A$, i.e., we have the partition of the process $\eta$ into $\eta \cap A$ and $\eta\setminus A$. We try to reconstruct the set $A$ just by the information contained in these two point sets. For that we approximate $A$ by the set $ A_\eta$ of all points in $\mathbb{R}^d$ which are closer to $\eta \cap A$ than to $\eta\setminus A$.
More formally, let $\eta$ be a homogeneous Poisson point process
of intensity ${\lambda} >0$, and denote by $\upsilon_\eta (x) = \{ z\in\mathbb{R}^d:\ \|
z-x \| \leqslant \| z-y \| \mbox{ for all } y \in \eta \} $ the
Voronoi cell generated by $\eta$ with nucleus $x\in\eta$. Then the
set $ A_\eta$ is just the union of the Poisson--Voronoi cells with
nuclei lying in $A$, i.e.,
$$ A_\eta = \bigcup\limits_{x \in \eta \cap A} \upsilon_\eta (x).$$
We call this set the {\it Poisson--Voronoi approximation } of the
set $A$. It was first introduced by Khmaladze and Toronjadze in~\cite{KhTo}. They proposed $A_\eta$ to be an estimator for $A$ when $\lambda$ is large (potential applications are listed in \cite[Section 1]{HR09}). They conjectured that for arbitrary bounded Borel set $A\subset\mathbb{R}^d, d\geqslant1,$ it holds
$$
\mbox{\rm Vol}(A_\eta)\to\mbox{\rm Vol}(A),\quad\lambda\to\infty,
$$
\begin{equation}\label{2047}
\mbox{\rm Vol}(A\Delta A_\eta)\to0,\quad\lambda\to\infty,
\end{equation}
almost surely, where $\mbox{\rm Vol}(\cdot)$ stands for the Lebesgue measure (volume) and $\Delta$ is the operation of the symmetric difference of sets. This conjecture was proved in~\cite{KhTo} for $d=1$. The case of general $d$ was treated by Einmahl and E. V. Khmaladze in \cite{EiKh} with some technical assumption on the boundary of $A$, and then generalized by Penrose in~\cite{Pen07} to an arbitrary bounded Borel set $A$.
It can be easily shown (see Section~\ref{sec:tools} for details) that for any Borel set $A$ it holds
$$
\mathbb{E}\, \mbox{\rm Vol}( A_\eta)= \mbox{\rm Vol}(A).
$$
Thus $\mbox{\rm Vol}( A_\eta)$ is an unbiased estimator for the
volume of $A$. In this paper we also consider the $n$--th moment of $\mbox{\rm Vol}( A_\eta)$ and approximate it by the $n$--th degree of the volume of the original set $\mbox{\rm Vol}^n( A)$ asymptotically as $\lambda\to\infty$ (Theorem~\ref{thm:asympt_volApprox}). For the case when $n=2$ and $A$ is a convex compact, similar estimates were obtained in \cite{HR09}.
It might be suggested from \eqref{2047} that
\begin{equation}\label{2140}
\mathbb{E}\,\mbox{\rm Vol}(A\Delta A_\eta)\to0,\quad\lambda\to\infty,
\end{equation}
although it is not a direct corollary. The more interesting problem is to find an exact asymptotic of $\mathbb{E}\,\mbox{\rm Vol}(A\Delta A_\eta)$. Initially it was considered by Heveling and Reitzner in \cite{HR09}. They proved that for any compact convex set $A$ with surface area $S(A)$ it holds
$$
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta)=c_d\cdot S(A)\cdot \lambda^{-1/d}(1+O(\lambda^{-1/d})),\quad\lambda\to\infty,
$$
where the constant $c_d$ independent of $\lambda$ and $A$ was calculated by them in an explicit form (see Section~\ref{sect:Results} for details).
Here we obtain a similar asymptotic formula (Theorem~\ref{thm:MeanVol}) for a much wider class of sets.
Namely, we consider Borel sets with finite volume $\mbox{\rm Vol} (A)$ and perimeter $\mathop{\rm Per}\nolimits(A)$ (see Section~\ref{sec:tools} for the precise definition). Our methods are completely different from those of Heveling and Reitzner. The key observations are the connection between the Poisson--Voronoi approximation and the covariogram of $A$, and the connection between the covariogram and the perimeter of a set recently established by Galerne \cite{Gal11}. As a by-product of our calculations, we prove that \eqref{2140} holds for any Borel set $A$ with finite volume (Corollary~\ref{cor:L1conv}).
We also consider higher moments of $\mbox{\rm Vol}(A \Delta A_\eta)$. For arbitrary Borel set $A$ we
approximate $\mathbb{E}\,\mbox{\rm Vol}^n(A \Delta A_\eta)$ by the $n$-th degree of $\mathbb{E}\,\mbox{\rm Vol}(A \Delta A_\eta)$ asymptotically
as $\lambda\to\infty$ (Theorem~\ref{0037}). Thus, assuming that $\mbox{\rm Vol} (A), \mathop{\rm Per}\nolimits(A)<\infty$ and using
the asymptotic for $\mathbb{E}\,\mbox{\rm Vol}(A \Delta A_\eta)$ from Theorem~\ref{thm:MeanVol}, we obtain the asymptotic for $\mathbb{E}\,\mbox{\rm Vol}^n(A \Delta A_\eta)$ (Corollary~\ref{cor:asympt_moments}).
\medskip
The paper is organized as follows. Our main results are stated in
the next section. In Section~\ref{sec:tools}, we introduce the
necessary background and notation, in particular the perimeter and
the covariogram of a set $A$. Proofs are given in Section
\ref{sect:Proofs}.
\section{Main results}\label{sect:Results}
Our first result yields the asymptotic of the average volume of $A\Delta A_\eta$ with increasing intensity $\lambda$. To formulate it, we need to define a notion of perimeter of a Borel set. The definition is somewhat technical, so we postpone it till Section~\ref{sec:tools}. If $A$ is a compact set with Lipschitz boundary (e.g. a convex body), then $\mathop{\rm Per}\nolimits(A)$ equals the $(d-1)$-dimensional Hausdorff measure
$\mathcal{H}_{d-1}(\partial A)$ of the boundary $\partial A$ of $A$. In general case it holds $\mathop{\rm Per}\nolimits(A)\leqslant\mathcal{H}_{d-1}(\partial A)$
(see, e.g. \cite[Proposition 3.62]{AF00}). Therefore, $\mathop{\rm Per}\nolimits(A)$ could be replaced by $\mathcal{H}_{d-1}(\partial A)$ in the assumptions of the theorem.
\begin{theorem}\label{thm:MeanVol}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$ and $\mathop{\rm Per}\nolimits(A)<\infty$, then
\begin{equation}\label{2340}
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta)=c_d\cdot \mathop{\rm Per}\nolimits(A)\cdot \lambda^{-1/d}(1+o(1)),\quad\lambda\to\infty,
\end{equation}
where $ c_d=2d^{-2}\Gamma(1/d)\kappa_{d-1}\kappa_d^{{-1-1/d}}$ and
$\kappa_n$ is the volume of the unit $n$-dimensional ball.
\end{theorem}
The probabilistic intuition behind this asymptotic is the
following. The set difference $A \Delta A_\eta$ behaves
asymptotically as a very small tube neighbourhood of the boundary
$\partial A$ formed out of the Poisson--Voronoi cells with nuclei
lying almost on $\partial A$. Since the volume of a typical
Poisson--Voronoi cell is $\lambda^{-1}$, its diameter has the
order $\lambda^{-1/d}$, and so the volume of this tube
neighborhood has the order $\mathop{\rm Per}\nolimits(A)\lambda^{-1/d}$.
\medskip
In the following, saying that some inequality holds asymptotically as $\lambda\to\infty$, we mean that it holds for
sufficiently large $\lambda\geqslant\lambda_0$. The choice of $\lambda_0$ might depend on $A$. Thus, all estimates are not uniform with
respect to $A$ (including those of Theorem~\ref{thm:MeanVol}).
\begin{theorem}\label{thm:asympt_volApprox}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$, then
$$
\Big|\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta)-\mbox{\rm Vol}^n(A)\Big|\leqslant C_{n,d}\cdot\mbox{\rm Vol}^{n-1}(A)\cdot\lambda^{-1},\quad\lambda\to\infty,
$$
where $C_{n,d}$ is some constant independent of $\lambda$ and $A$.
\end{theorem}
\begin{remark}
In fact, we show that the following non--asymptotic inequality holds: for any $\lambda>0$
$$
\Big|\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta)-\mbox{\rm Vol}^n(A)\Big|\leqslant C_{n,d}\cdot\sum_{k=1}^{n-1}\mbox{\rm Vol}^{n-k}(A)\cdot\lambda^{-k}.
$$
\end{remark}
\begin{theorem}\label{0037}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$ and $\mathop{\rm Per}\nolimits(A)<\infty$, then
$$
\Big|\mathbb{E}\,\mbox{\rm Vol}^n(A\Delta A_\eta)-(\mathbb{E} \mbox{\rm Vol}(A\Delta A_\eta))^n\Big|\leqslant C^\prime_{n,d} \cdot \mathop{\rm Per}\nolimits (A)^{n-1} \cdot\lambda^{-1-(n-1)/d},\quad\lambda\to\infty,
$$
where $C^\prime_{n,d}$ is some constant independent of $\lambda$ and $A$.
\end{theorem}
\begin{remark}
We conjecture that the following limit theorems can be proven by
the method of moments (see e.g. {\rm \cite[Theorems 30.1,
30.2]{Bill79}}):
\begin{equation} \label{eq:CLT}
\lambda^{1/2(1+1/d)}\left( \mbox{\rm Vol}(A_\eta) - \mbox{\rm Vol}(A) \right) \to
N(0,\sigma_1 \mathop{\rm Per}\nolimits(A)),
\end{equation}
$$
\lambda^{1/2(1+1/d)}\left( \mbox{\rm Vol}(A \Delta A_\eta) - c_d \mathop{\rm Per}\nolimits(A)
\lambda^{-1/d} \right) \to N(0,\sigma_2 \mathop{\rm Per}\nolimits(A) )
$$
in distribution as $\lambda\to\infty,$ $\sigma_1, \sigma_2>0$.
\end{remark}
Recently (\ref{eq:CLT}) was proved by Schulte~\cite{Schulte2012} for {\it convex} sets $A$ using a central limit theorem for Wiener-It\'o chaos expansions. In his Remark~4 he points out that the result can be extended to all sets where the volume of a small tube neighbourhood $B(\partial A)$ of $\partial A$ can be bounded in a nice way. Yet the general conjecture seems to be open.
\begin{corollary}\label{cor:asympt_moments}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$ and $\mathop{\rm Per}\nolimits(A)<\infty$, then
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A\Delta A_\eta)=\left( \mathbb{E} \mbox{\rm Vol}(A\Delta
A_\eta)\right)^n(1+O(\lambda^{-1+1/d})),\quad\lambda\to\infty,
$$
and for $d\geqslant2$
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A\Delta A_\eta)=(c_d\mathop{\rm Per}\nolimits(A))^n\lambda^{-n/d}(1+o(1)),\quad\lambda\to\infty.
$$
\end{corollary}
The asymptotic order of the variance of $A_\eta$ and $A \Delta
A_\eta$ as $\lambda\to\infty$ was first studied in \cite{HR09} for
convex sets $A$. We extend that results to arbitrary Borel sets.
\begin{corollary} \label{cor:var}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$ and $\mathop{\rm Per}\nolimits(A)<\infty$, then
$$
\mbox{\rm Var}\, \mbox{\rm Vol}(A_\eta)\leqslant C_d\cdot \mathop{\rm Per}\nolimits(A)\cdot \lambda^{-1- 1/d},\quad\lambda\to\infty,
$$
and
$$
\mbox{\rm Var}\, \mbox{\rm Vol}(A \Delta A_\eta)\leqslant C_d\cdot \mathop{\rm Per}\nolimits(A)\cdot \lambda^{-1- 1/d},\quad\lambda\to\infty,
$$
where $C_{d}$ is some constant independent of $\lambda$ and $A$.
\end{corollary}
The second inequality follows immediately from Theorem~\ref{0037}. The first inequality will be proved in Section~\ref{sect:Moments}.
\medskip
The probabilistic heuristic explaining the asymptotic behavior of
the variances is the following. Since $A \Delta A_\eta$ is
asymptotically a very small tube neighbourhood $B(\partial A)$ of
$\partial A$ consisting of parts $\tilde \upsilon_\eta(x)$ of almost independent Poisson--Voronoi cells $\upsilon_\eta(x)$
with nuclei $x\in B(\partial A)$ we may use the formula for the
variance of the compound Poisson distribution:
$$
\mbox{\rm Var}\, \mbox{\rm Vol}(A \Delta A_\eta) = \mbox{\rm Var}\, \left( \sum_{x\in \eta \cap B(\partial A)}
\mbox{\rm Vol}(\tilde \upsilon_\eta(x)) \right)\approx \mbox{\rm Var}\, \left( \sum_{i=1}^N Y_i
\right)
$$
where random variables $Y_i\stackrel{d}{=} \mbox{\rm Vol}(\tilde \upsilon_\eta(x))$ are {\it
i.i.d.} and $$N\stackrel{d}{=} {\rm card} (\eta \cap B(\partial A) )\sim
\mbox{Pois}({\lambda} \mbox{\rm Vol}(B(\partial A)))$$ is independent of $Y_i$. Here
$\stackrel{d}{=}$ means the equality in distribution and ${\rm card} (B)$ is
the cardinality of a set $B$. Then
\begin{multline*}
\mbox{\rm Var}\, \left( \sum_{i=1}^N Y_i
\right) =\mathbb{E}\, N \, \mbox{\rm Var}\, Y_1 + \mbox{\rm Var}\, N \, (\mathbb{E}\, Y_1)^2={\lambda}
\mbox{\rm Vol}(B(\partial A))\, \mathbb{E}\, Y_1^2\\
\le \lambda \mbox{\rm Vol} (B(\partial A)) \left( \mathbb{E}\, \mbox{\rm Vol} (\upsilon_\eta(x)) \right)^2 =O\left({\lambda} \mathop{\rm Per}\nolimits(A) {\lambda}^{-1/d}
{\lambda}^{-2}\right)=\mathop{\rm Per}\nolimits(A) \, O\left({\lambda}^{-1-1/d}\right)
\end{multline*} since $\tilde \upsilon_\eta(x)\subset \upsilon_\eta(x)$ for any $x$, the
second moment of the volume of a typical Poisson--Voronoi cell is
of order $\lambda^{-2}$ and the volume of $B(\partial A)$ is of
order $\mathop{\rm Per}\nolimits(A)\lambda^{-1/d}$.
\bigskip
The results of Corollary $\ref{cor:var}$ can also be obtained by using
the Poincar\'e inequality which gives
an upper bound on the variance of a functional of a
Poisson point process. Let $\cal N$ be the set of all locally finite configurations on $\mathbb{R}^d$. Consider a nonnegative measurable function $F\,:\, \cal N\to\mathbb{R}$. If $\mathbb{E}\, F^2(\eta) < \infty$, then
\begin{equation}\label{eq:EfronStein}
\mbox{\rm Var}\, F(\eta ) \leqslant {\lambda} \mathbb{E}\, \int\limits_{\mathbb{R}^{d}} ( F(\eta \cup \{y\}
)- F(\eta))^2 \, dy,
\end{equation}
where we added a point $y$ to the Poisson point process $\eta$. Putting $F(\eta)=\mbox{\rm Vol}(A_\eta)$ in \eqref{eq:EfronStein}, we get
$$
\mbox{\rm Var}\, \mbox{\rm Vol}(A_\eta) \leqslant {\lambda} \int\limits_{\mathbb{R}^{d}} \mathbb{E}\, \left(
\mbox{\rm Vol}(A_{\eta\cup \{y\} }) - \mbox{\rm Vol}(A_\eta) \right)^2 \, dy,
$$
where the right--hand side can be estimated from above to get the
upper bound in Corollary $\ref{cor:var}$. The reasoning for the symmetric difference $A\Delta A_\eta$ is similar.
In full generality, inequality \eqref{eq:EfronStein} was proved by Wu {\rm \cite{Wu}.}
As was shown by Last and Penrose {\rm \cite[Theorem 1.2]{LastPenrose2010}}, it
is a consequence of an even more general inequality following from
the Fock space representation of Poisson point processes.
\section{Preliminaries}\label{sec:tools}
For basic facts from
integral geometry, stochastic geometry and Voronoi tessellations
which are not explained in the following, we refer the reader to
\cite{SchnWe3}, \cite{SKM95}, and \cite{Mol}.
Define the perimeter of a Borel set $A$ as
$$
\mathop{\rm Per}\nolimits(A)=\sup\Big\{\int\limits_{A}\mathop{\rm div}\nolimits\varphi(x)\, dx\,:\,\varphi\in\mathcal{C}_c^1(\mathbb{R}^d),\|\varphi\|_\infty\leqslant1\Big\},
$$
cf. \cite{AF00}, where
$$
\mathop{\rm div}\nolimits\varphi(x)=\sum_{i=1}^d\frac{\partial\varphi_i}{\partial
x_i}\quad\mbox{and}\quad\|\varphi\|_\infty=\max_{i=1,\dots,d}\sup_{x\in\mathbb{R}^d}|\varphi_i(x)|
$$
for $\varphi=(\varphi_1,\dots,\varphi_d)$. The class
$\mathcal{C}_c^1(\mathbb{R}^d)$ consists of all continuously
differentiable vector--valued functions from $\mathbb{R}^d$ to
$\mathbb{R}^d$ with compact support.
Let $A$ be a Borel set with finite volume. Then
$$
g_A(x)=\mbox{\rm Vol}((A+x)\cap A),\quad x\in\mathbb{R}^d,
$$
is a covariogram of $A$. For the history on the covariogram
problem see the references in \cite{Gal11} and also the recent
breakthrough by Averkov and Bianchi \cite{AvBi09}.
In the proof of Theorem~\ref{thm:MeanVol} we use the result obtained by Galerne in~\cite[Theorem~14]{Gal11}. The following assertions are equivalent:
\begin{itemize}
\item[(a)] $\mathop{\rm Per}\nolimits(A)<\infty$;
\item[(b)] there exists a finite limit
\begin{equation}\label{eq:1}
\lim_{r\to+0}\frac{g_A(ru)-g_A(0)}{r} = \frac{\partial g_A}{\partial u}(0)
\end{equation}
for all $u\in\mathbb{S}^{d-1}$;
\item[(c)] $g_A$ is Lipschitz.
\end{itemize}
In addition, the Lipschitz constant of $g_A$ satisfies
\begin{equation}\label{0416}
\mbox{\rm Lip}(g_A)\leqslant\frac12\mathop{\rm Per}\nolimits(A)
\end{equation}
and it holds
\begin{equation}\label{eq:2}
\int\limits_{\mathbb{S}^{d-1}}\frac{\partial g_A}{\partial u}(0) \, {\mathcal H}_{d-1} (du) = - \kappa_{d-1} \mathop{\rm Per}\nolimits(A).
\end{equation}
\bigskip
Another tool we need is the refined Campbell--Mecke formula for stationary point processes (cf. e.g. \cite{SKM95}). Using Slivnyak's theorem, we give its particular case for the Poisson point process.
As above, let $\eta$ be a homogeneous Poisson point process of intensity ${\lambda} >0$, and $\cal N$ be the set
of all locally finite point configurations on $\mathbb{R}^d$. Consider a nonnegative measurable function $f\,:\, {\cal N}\times(\mathbb{R}^d)^m\to\mathbb{R}$. Then
\begin{equation}\label{2352}
\mathbb{E}\,\sum_{(y_1,\dots,y_m)\in\eta^m_{\ne}}F(\eta,y_1,\dots,y_m)=
\lambda^m\int\limits_{(\mathbb{R}^d)^m}\mathbb{E}\, F(\eta\cup{\bar y_{m}},y_1,\dots,y_m)\,dy_1\dots dy_m,
\end{equation}
where $\eta^m_{\ne}$ denotes the set of all $m$--tuples of pair--wise distinct points from $\eta$,
and $\eta\cup{\bar y_{m}}$ is the process $\eta$ with added point set $\bar y_{m} =\{ y_1,\dots, y_m \}$.
As a simple corollary we get two identities which are crucial for us in the sequel.
\begin{proposition}
If $A \subset \mathbb{R}^d $ is a Borel set with $\mbox{\rm Vol}(A)<\infty$, then
\begin{equation}\label{1410}
\mathbb{E}\mbox{\rm Vol}(A_\eta)=\lambda\int\limits_{\mathbb{R}^d}\int\limits_{A}e^{-\lambda\kappa_d\|y-x\|^d}\,dy\,dx=\mbox{\rm Vol}(A),
\end{equation}
\begin{equation}\label{1411}
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta)=
2\lambda \int\limits_{\mathbb{R}^d\setminus A} \int\limits_{A} e^{-\lambda\kappa_d\|y-x\|^d} \,dy \, dx.
\end{equation}
\end{proposition}
\begin{proof}
By Fubini's theorem and the Slivnyak-Mecke formula \eqref{2352}, we have
\begin{multline*}
\mathbb{E}\,\mbox{\rm Vol}(A_\eta)=
\mathbb{E}\, \int\limits_{\mathbb{R}^d}{\bf 1} \left(x\in A_\eta\right)\,dx=\int\limits_{\mathbb{R}^d}\mathbb{E}\,\sum_{y\in\eta\cap A}{\bf 1} \left(x\in\upsilon_{\eta}(y)\right)\,dx
\\=\lambda\int\limits_{\mathbb{R}^d} \int\limits_{A}\P\left(x\in\upsilon_{\eta\cup\{y\}}(y)\right) \,dy \,dx
=\lambda\int\limits_{\mathbb{R}^d} \int\limits_{A}e^{-\lambda\kappa_d\|x-y\|^d} \,dy \,dx .
\end{multline*}
Similarly, we obtain
$$
\mathbb{E}\,\mbox{\rm Vol}(A\setminus A_\eta)=\lambda\int\limits_{A} \int\limits_{\mathbb{R}^d\setminus A}e^{-\lambda\kappa_d\|x-y\|^d}\,dy \,dx
$$
and
$$
\mathbb{E}\,\mbox{\rm Vol}(A_\eta\setminus A)=\lambda\int\limits_{\mathbb{R}^d\setminus A} \int\limits_{A}e^{-\lambda\kappa_d\|x-y\|^d}\,dy \,dx .
$$
By definition $\mbox{\rm Vol}(A\Delta A_\eta)=\mbox{\rm Vol}(A\setminus A_\eta)+\mbox{\rm Vol}(A_\eta\setminus A)$ which completes the proof of \eqref{1411}. To prove the second part of \eqref{1410}, one has to apply Fubini's theorem and then use the formula
\begin{equation}\label{0609}
\int\limits_{\mathbb{R}^d}e^{-c\|x-y\|^d}\,dx =\frac{\kappa_d}c,\quad c>0,
\end{equation}
which could be easily proved by introducing spherical coordinates.
\end{proof}
Notice that we have also proved that
\begin{equation*}
\mathbb{E}\,\mbox{\rm Vol}(A\setminus A_\eta)=\mathbb{E}\,\mbox{\rm Vol}(A_\eta\setminus A).
\end{equation*}
However, $\mbox{\rm Vol}(A\setminus A_\eta)$ and $\mbox{\rm Vol}(A_\eta\setminus A)$ are not equidistributed since the first random variable is bounded, and the second is not.
As a direct colollary of the identity~(\ref{1410}) we get
\begin{equation}\label{1543}
\mbox{\rm Var}\,\mbox{\rm Vol}( A_\eta)=\mathbb{E}\,\left( \mbox{\rm Vol}(A\setminus A_\eta)-\mbox{\rm Vol}(A_\eta\setminus A)\right)^2,
\end{equation}
which we shall use in the following.
\section{Proofs}\label{sect:Proofs}
\subsection{Asymptotics of the mean volume of the symmetric
difference }\label{sect:MeanVol}
In this section we give the proof of Theorem \ref{thm:MeanVol}. The key step to prove it is the
following relation between the Poisson--Voronoi approximation and
the covariogram of a set $A$.
\begin{lemma}\label{lemma:meanV_d}
Let $g_A(x)$ be the covariogram of a Borel set $A$ with
$\mbox{\rm Vol}(A)<\infty$. Then
\begin{equation} \label{eq:symmDiff}
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta) = -2 \int\limits_{0}^\infty r^{d-1}
e^{- \kappa_dr^d}\,
\tilde{g}_A( {\lambda}^{- 1/d} r ) \, dr \, ,
\end{equation}
where
\begin{equation}\label{def:Sg}
\tilde{g}_A (r) = \int\limits_{\mathbb{S}^{d-1}} \left(g_A(ru) -
g_A(0)\right) \, {\mathcal H}_{d-1} (du).
\end{equation}
\end{lemma}
\begin{proof}
Replacing $y$ in \eqref{1411} by $x-{\lambda}^{-1/d} z$ we get
\begin{eqnarray*}
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta) &=&
2\lambda\int\limits_{\mathbb{R}^d}\int\limits_{\mathbb{R}^d}e^{-\lambda\kappa_d\|y-x\|^d}\mathbf{1}\{y\in
A,x\in A^c\} \, dy \, dx
\\ &=&
2 \int\limits_{\mathbb{R}^d}e^{- \kappa_d\|z\|^d}\int\limits_{\mathbb{R}^d}\mathbf{1}\{x\in(A+ {\lambda}^{-1/d} z)\cap A^c\} \,dz \,dx
\\ &=&
2 \int\limits_{\mathbb{R}^d}e^{- \kappa_d\|z\|^d} \mbox{\rm Vol}((A+{\lambda}^{-1/d} z)\cap A^c) \, dz.
\end{eqnarray*}
By the definition of the covariogram
$ \mbox{\rm Vol}((A+ {\lambda}^{-1/d} z)\cap A^c)=g_A(0)-g_A({\lambda}^{-1/d} z) $.
We introduce spherical coordinates $z=ru$, where
$r\in\mathbb{R}^+$ and $u\in\mathbb{S}^{d-1}$. This yields
$$
\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta) = -2 \int\limits_{0}^\infty r^{d-1}
e^{- \kappa_dr^d}\, \Big[ \int\limits_{{\bf S}^{d-1}}\left( g_A({\lambda}^{- 1/d}
ru) - g_A(0) \right) \, {\mathcal H}_{d-1} (du) \Big]\, dr \, .
$$
\end{proof}
\begin{corollary}\label{cor:L1conv}
For any measurable $A$ with $\mbox{\rm Vol}(A)<\infty$ it holds
$$
\mathbb{E}\mbox{\rm Vol}(A\Delta A_\eta)\to 0,\quad\lambda\to\infty.
$$
\end{corollary}
\begin{proof}
It immediately follows from \eqref{eq:symmDiff} and the
continuity of the set covariogram.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm:MeanVol}]
Using Lemma \ref{lemma:meanV_d} and substituting $t$ for $\kappa_d r^d$ we obtain
$$
\mathbb{E}\mbox{\rm Vol}(A\Delta A_\eta) = -\frac2{d\kappa_d}
\int\limits_{0}^\infty e^{-t}\, \tilde{g}_A \left((\lambda\kappa_d)^{-1/d}
t^{1/d}\right) \, dt\,.
$$
It follows from \eqref{0416} and the definition of $\tilde{g}_A$ that
\begin{equation}\label{eq:SgLip}
|\tilde{g}_A (r)| \leqslant \frac 12 {\mathcal H}_{d-1} ({\bf S}^{d-1}) \mathop{\rm Per}\nolimits(A) r .
\end{equation}
Therefore, Lebesgue's Dominated Convergence Theorem and equations \eqref{eq:1}, \eqref{eq:2} yield
\begin{multline*}
\lim_{\lambda\to\infty} \mathbb{E}\mbox{\rm Vol}(A\Delta A_\eta) {\lambda}^{1/d}
= -\frac2 d \kappa_d^{-1-1/d} \lim_{\lambda\to\infty}
\int\limits_{0}^\infty e^{-t} t^{1/d} \, \frac {\tilde{g}_A
((\lambda\kappa_d)^{-1/d} t^{1/d})
}{(\lambda\kappa_d)^{-1/d} t^{1/d}} \, dt
\\= -\frac2 d \kappa_d^{-1-1/d}\int\limits_{0}^\infty e^{-t} t^{1/d}dt\int\limits_{{\bf S}^{d-1}} \frac{\partial g_A}{\partial u}(0)\, {\mathcal H}_{d-1} (du)
\\=\frac2 d \kappa_{d-1} \kappa_d^{-1-1/d} \mathop{\rm Per}\nolimits (A)
\int\limits_{0}^\infty e^{-t} t^{1/d} \, dt
=\frac2 d \kappa_{d-1} \kappa_d^{-1-1/d}\Gamma\Big(1+\frac1d\Big) \mathop{\rm Per}\nolimits (A).
\end{multline*}
\end{proof}
\subsection{Asymptotics of higher moments} \label{sect:Moments}
To prove Theorem~\ref{thm:asympt_volApprox} and Theorem \ref{0037}, we need a number of lemmas. In this section $C$ is always some constant independent of $\lambda$ and $A$. Our first statement is the following version of H\"older's inequality.
\begin{lemma}\label{2301}
For any events $A_1,\dots,A_m$ it holds
$$
\P\left(\bigcap_{r=1}^mA_r\right)\leqslant\prod_{r=1}^m
\left(\P(A_r)\right)^{1/m}.
$$
\end{lemma}
\begin{lemma}\label{1534}
Let $x_0,y_0\in\mathbb{R}^d$. For any $\varepsilon>0$ and $m\in\mathbb{N}$ the
following inequality holds:
$$
\int\limits_{(\mathbb{R}^d)^m} \left( \P (x_0,x_1,\dots,
x_m\in\upsilon_{\eta\cup\{y_0\}}(y_0))\right)^\varepsilon
\, dx_1\dots dx_m \leqslant
e^{-\varepsilon\lambda\kappa_d\|x_0-y_0\|^d/(m+1)}\left(\frac{m+1}{\varepsilon\lambda}\right)^m.
$$
\end{lemma}
\begin{proof} By Lemma \ref{2301}, we have
\begin{multline*}
\int\limits_{(\mathbb{R}^d)^m}\left( \P (x_0,x_1,\dots,
x_m\in\upsilon_{\eta\cup\{y_0\}}(y_0))\right)^\varepsilon
\, dx_1\dots dx_m
\\
\leqslant\left( \P(x_0\in\upsilon_{\eta\cup\{y_0\}}(y_0)) \right)^{\varepsilon/(m+1)}
\int\limits_{(\mathbb{R}^d)^m} \prod_{i=1}^m\left( \P (x_i\in\upsilon_{\eta\cup\{y_0\}}(y_0)) \right)^{\varepsilon/(m+1)}
\, dx_1\dots dx_m
\\
=e^{-\varepsilon\lambda\kappa_d\|x_0-y_0\|^d/(m+1)}
\left[\ \int\limits_{\mathbb{R}^d} \, e^{-\varepsilon \lambda\kappa_d \|x-y_0\|^d/(m+1)} \, dx \right]^m.
\end{multline*}
Using \eqref{0609} completes the proof.
\end{proof}
\begin{lemma}\label{0623}
For any $a>0$
\begin{equation}\label{0628}
\int\limits_{\mathbb{R}^d} \int\limits_{A}e^{-a\lambda\|y-x\|^d}\,dy \,dx = \frac{\kappa_d \mbox{\rm Vol}(A)}{a \lambda}
\end{equation}
and
\begin{equation}\label{0629}
\int\limits_{\mathbb{R}^d\setminus A} \int\limits_{A}e^{-a\lambda\|y-x\|^d}\,dy \,dx
\leqslant C\frac{\mathop{\rm Per}\nolimits(A)}{\lambda^{1+1/d}},\quad \lambda\to\infty.
\end{equation}
\end{lemma}
\begin{proof}
The first equation follows from \eqref{1410} after replacing $\lambda$ by $\lambda' a/ \kappa_d$. The second estimate
follows from \eqref{1411} after replacing $\lambda$ by $\lambda' a/ \kappa_d$ and then applying Theorem~\ref{thm:MeanVol}.
\end{proof}
Introduce the notation $B_r^x$ for the closed ball with center $x\in\mathbb{R}^{d}$ and radius $r>0$ in Euclidean metric.
\begin{lemma}\label{1654}
Let $x_1,x_2,y_1,y_2\in\mathbb{R}^d$. If $B^{x_1}_{\|x_1-y_1\|}\cap B^{x_2}_{\|x_2-y_2\|}\ne\emptyset$, then
$$
\P\left(B^{x_1}_{\|x_1-y_1\|}\cap\eta=\emptyset,B^{x_2}_{\|x_2-y_2\|}\cap\eta=\emptyset\right)
\leqslant2\exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}}\left(\|x_1-y_2\|^d+\|x_2-y_1\|^d\right)\right).
$$
\end{lemma}
\begin{proof}
Since $B^{x_1}_{\|x_1-y_1\|}\cap B^{x_2}_{\|x_2-y_2\|}\ne\emptyset$, it follows from the triangle inequality that
$$
\frac{\|x_1-y_2\|}{4},\frac{\|x_2-y_1\|}{4}\leqslant\max\left(\|x_1-y_1\|,\|x_2-y_2\|\right).
$$
Therefore, by Lemma \ref{2301} and stationarity of $\eta$ we have
\begin{multline*}
\P\left(B^{x_1}_{\|x_1-y_1\|}\cap\eta=\emptyset,B^{x_2}_{\|x_2-y_2\|}\cap\eta=\emptyset\right)
\\\leqslant\P\left(B^{x_1}_{\|x_1-y_2\|/4}\cap\eta=\emptyset,B^{x_1}_{\|x_2-y_1\|/4}\cap\eta=\emptyset\,\text{or}\,B^{x_2}_{\|x_1-y_2\|/4}\cap\eta=\emptyset,B^{x_2}_{\|x_2-y_1\|/4}\cap\eta=\emptyset\right)
\\\leqslant \sum_{i=1}^2 \P\left(B^{x_i}_{\|x_1-y_2\|/4}\cap\eta=\emptyset,B^{x_i}_{\|x_2-y_1\|/4}\cap\eta=\emptyset\right)
\\\leqslant \sum_{i=1}^2 \left( \P\left(B^{x_i}_{\|x_1-y_2\|/4}\cap\eta=\emptyset\right)\P\left(B^{x_i}_{\|x_2-y_1\|/4}\cap\eta=\emptyset\right)\right)^{1/2}
\\=2\exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}}\left(\|x_1-y_2\|^d+\|x_2-y_1\|^d\right)\right).
\end{multline*}
\end{proof}
\begin{lemma}\label{1952}
For any $x_1,y_1,\dots,x_n,y_n\in\mathbb{R}^d$ it holds
\begin{multline*}
\P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset,r=1,\dots,n\right)\leqslant \exp\left(-\lambda\kappa_d\sum_{r=1}^n\|x_r-y_r\|^d\right)
\\+2\sum_{s<t}\exp\left(-\frac{\lambda\kappa_d}{n+1}\sum_{r=1}^n\|x_r-y_r\|^d\right)\exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right).
\end{multline*}
\end{lemma}
\begin{proof}
If the balls $B^{x_r}_{\|x_r-y_r\|},\,r=1,\dots,n$ are pairwise disjoint then we obviously have
$$
\P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset,r=1,\dots,n\right)=\exp\left(-\lambda\kappa_d\sum_{r=1}^n\|x_r-y_r\|^d\right).
$$
Suppose that for some indices $s\ne t$ it holds $B^{x_s}_{\|x_s-y_s\|}\cap B^{x_t}_{\|x_t-y_t\|}\ne\emptyset$. Applying Lemma~\ref{2301}, we get
\begin{multline*}
\P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset,r=1,\dots,n\right)
\\\leqslant \left( \P\left(B^{x_s}_{\|x_s-y_s\|}\cap\eta=\emptyset,\,B^{x_t}_{\|x_t-y_t\|}\cap\eta=\emptyset\right)\right)^{1/(n+1)} \prod_{r=1}^n \left( \P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset\right) \right)^{1/(n+1)}
\\=\exp\left(-\frac{\lambda\kappa_d}{n+1}\sum_{r=1}^n\|x_r-y_r\|^d\right)\left( \P\left(B^{x_s}_{\|x_s-y_s\|}\cap\eta=\emptyset,\,B^{x_t}_{\|x_t-y_t\|}\cap\eta=\emptyset\right)\right)^{1/(n+1)}.
\end{multline*}
It remains to apply Lemma~\ref{1654} to finish the proof.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:asympt_volApprox}]
We have
\begin{multline}\label{0714}
\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta) =
\mathbb{E}\,\int\limits_{(\mathbb{R}^d)^{n}} {\bf 1} ( \exists (y_1,\dots, y_n)\in (\eta\cap A)^n\,:\,x_i\in\upsilon_\eta(y_i), i=1,\dots n) \, dx_{1}\dots dx_{n}
\\=\sum_{i=1}^n\sum_{m_1+\dots+m_i=n}B_{n,i,m_1,\ldots,m_i}\beta_{i,m_1,\ldots,m_i},
\end{multline}
where
\begin{multline*}
\beta_{i,m_1,\ldots,m_i}=\int\limits_{(\mathbb{R}^d)^{n}}
\mathbb{E}\, \! \sum_{(y_1,\dots,y_i)\in(\eta\cap A)^i_{\ne}} \!\!
{\bf 1} \left( x_1,\dots,x_{m_1}\in\upsilon_\eta(y_1), \dots ,x_{n-m_i+1},\dots,x_n\in\upsilon_\eta(y_i)\right)
\\ \, dx_{1}\dots dx_{n}
\end{multline*}
and $B_{n,i,m_1,\ldots,m_i}$ denotes the number of ways to divide the set $\{1,2,\dots,n\}$ into $i$ subsets of size $m_1,\ldots,m_i$. It it clear that
\begin{equation}\label{0942}
B_{n,n,1,\dots,1}=1.
\end{equation}
Fix some $i$ and $m_1,\dots,m_i$. Using the Slivnyak-Mecke formula \eqref{2352} we
get
\begin{multline*}
\beta_{i,m_1,\ldots,m_i}=\lambda^{i}\int\limits_{(\mathbb{R}^d)^n}\int\limits_{A^i}
\P \left( x_1,\dots,x_{m_1}\in\upsilon_{\eta\cup \tilde{y}_{i} }(y_1),\dots, x_{n-m_i+1},\dots,x_n\in\upsilon_{\eta\cup \tilde{y}_{i}}(y_i)\right)
\\ dy_1\dots dy_i \, dx_1\dots dx_n ,
\end{multline*}
where $\tilde{y}_{i}=\{ y_1,\ldots, y_{i} \}$. Taking into account that
$\upsilon_{\eta\cup \tilde{y}_{i} }(y_r) \subset \upsilon_{\eta\cup \{y_r\} }(y_r) $,
and using Fubini's theorem, Lemma~\ref{2301} and Lemma~\ref{1534} we obtain
\begin{eqnarray*
\beta_{i,m_1,\ldots,m_i}
&\leqslant&
\lambda^{i}\int\limits_{A^i} \prod_{r=1}^i\int\limits_{(\mathbb{R}^d)^{m_r}}
\left( \P\left( x_1,\dots,x_{m_r}\in\upsilon_{\eta\cup\{y_r\}}(y_r) \right)\right)^{1/i}
\, dx_1\dots dx_{m_r} \, dy_1\dots dy_i
\\ &\leqslant&
\lambda^{i}\int\limits_{A^i} \prod_{r=1}^i \Big( \frac{im_r}{ {\lambda}} \Big)^{m_r-1}
\int\limits_{\mathbb{R}^d} \Big( e^{- \frac 1i {\lambda} \kappa_d \| x_1-y_r\|^d /m_r} \Big) \, dx_1 \, dy_1\dots dy_i .
\end{eqnarray*}
By \eqref{0628} we get
$$
\beta_{i,m_1,\ldots,m_i}\leqslant
C\mbox{\rm Vol}^i(A)\lambda^{i-\sum_{r=1}^im_r}=C\mbox{\rm Vol}^i(A)\lambda^{i-n}.
$$
The maximum order of $\lambda$ is achieved for $i=n$, which together with \eqref{0714} and \eqref{0942} implies
\begin{multline*}
\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta)
\leqslant
\lambda^{n} \int\limits_{(\mathbb{R}^d)^n} \int\limits_{A^n}
\P\left( x_r\in\upsilon_{\eta\cup \tilde{y}_{n}}(y_r),\; r=1,\dots,n\right)
\, dy_1\dots dy_n \, dx_1\dots dx_n
\\ +
C(\mbox{\rm Vol}(A))^{n-1}\lambda^{-1},\quad\lambda\to\infty.
\end{multline*}
It is clear that
$$
\P\left( x_r\in\upsilon_{\eta\cup \tilde{y}_{n}}(y_r),\; r=1,\dots,n\right)\leqslant\P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset,r=1,\dots,n\right).
$$
Therefore, by Lemma~\ref{1952},
\begin{equation}\label{1048}
\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta) \leqslant v_n+2\sum_{s<t}v_{n,s,t}+C(\mbox{\rm Vol}(A))^{n-1}\lambda^{-1},\quad\lambda\to\infty,
\end{equation}
where
$$
v_n=\lambda^{n}\int\limits_{(\mathbb{R}^d)^n} \int\limits_{A^n} \exp\left(-\lambda\kappa_d\sum_{r=1}^n\|x_r-y_r\|^d\right)
\, dy_1\dots dy_n \, dx_1\dots dx_n,
$$
and
\begin{multline*}
v_{n,s,t}=\lambda^{n}\int\limits_{(\mathbb{R}^d)^n} \int\limits_{A^n}
\exp\left(-\frac{\lambda\kappa_d}{n+1}\sum_{r=1}^n\|x_r-y_r\|^d\right)
\\ \times \exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right)
\, dy_1\dots dy_n \, dx_1\dots dx_n .
\end{multline*}
By formula~\eqref{1410},
\begin{equation}\label{1049}
v_n=\mbox{\rm Vol}^n(A).
\end{equation}
Let us estimate $v_{n,s,t}$. Using Fubini, it follows from \eqref{0628} that
\begin{eqnarray*}
v_{n,s,t}
&\leqslant& C\mbox{\rm Vol}^{n-2}(A)\lambda^{2}
\int\limits_{\mathbb{R}^d} \int\limits_{\mathbb{R}^d} \int\limits_{A} \int\limits_{A}
\exp\left(-\frac{\lambda\kappa_d}{(n+1)}\left(\|x_s-y_s\|^d+\|x_t-y_t\|^d\right) \right)
\\ &&
\hskip2cm \times\exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right)
\, dy_t \, dy_s \, dx_t \, dx_s
\\
&\leqslant& C\mbox{\rm Vol}^{n-2}(A)\lambda^{2}
\int\limits_{\mathbb{R}^d} \int\limits_{A}
\exp\left(-\frac{\lambda\kappa_d}{(n+1)}\left(\|x_s-y_s\|^d\right) \right)
\\ &&
\hskip1.2cm \times \int\limits_{\mathbb{R}^d} \int\limits_{\mathbb{R}^d} \exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right)
\, dy_t \, dx_t \, dy_s \, dx_s .
\end{eqnarray*}
Furthermore, by \eqref{0609},
$$
v_{n,s,t}\leqslant C\mbox{\rm Vol}^{n-2}(A)
\int\limits_{\mathbb{R}^d} \int\limits_{A} \exp\left(-\frac{\lambda\kappa_d}{(n+1)}\|x_s-y_s\|^d\right) \, dy_s \, dx_s ,
$$
and applying \eqref{0628} again, we get
$$
v_{n,s,t}\leqslant C\mbox{\rm Vol}^{n-1}(A)\lambda^{-1}.
$$
Combining this with the estimate~\eqref{1048} and with \eqref{1049}, we
get
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta) \leqslant \mbox{\rm Vol}^n(A)+C\mbox{\rm Vol}^{n-1}(A)\lambda^{-1}, \quad \lambda \to \infty.
$$
The application of Lyapunov's inequality
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A_\eta)\geqslant (\mathbb{E}\,\mbox{\rm Vol}(A_\eta))^n= \mbox{\rm Vol}^n(A)
$$
finishes the proof.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{0037}]
We have
\begin{equation}\label{0032}
\mathbb{E}\, \mbox{\rm Vol}^n(A\Delta A_\eta)=\mathbb{E}\,\left( \mbox{\rm Vol}(A\setminus
A_\eta)+\mbox{\rm Vol}(A_\eta\setminus A)\right)^n
\\=\sum_{k=0}^n \binom n k u_k,
\end{equation}
where
\begin{equation}\label{1516}
u_k = \mathbb{E}\, \int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k}
{\bf 1} (x_1,\dots x_k\in A_\eta,x_{k+1},\dots,x_n\not\in A_\eta)
\, dx_1\dots dx_n.
\end{equation}
Fix some $k$. We have
\begin{multline}\label{0736}
u_k = \mathbb{E}\, \int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k}
{\bf 1} \Big( \exists (y_1,\dots, y_k)\in (\eta\cap A)^k, (y_{k+1},\dots, y_n)\in(\eta\setminus A)^{n-k}\,:
\\ \hfill x_i\in\upsilon_\eta(y_i), i=1,\dots n \Big)
\, dx_1\dots dx_n
\\=\sum_{i=1}^k\sum_{j=1}^{n-k}\sum_{m_1+\dots+m_i=k}\sum_{l_1+\dots+l_j=n-k}B_{k,i,m_1,\ldots,m_i}B_{n-k,j,l_1,\ldots,l_j}\beta_{i,j,m_1,\ldots,m_i,l_1,\ldots,l_j},
\end{multline}
where
\begin{multline*}
\beta_{i,j,m_1,\ldots,m_i,l_1,\ldots,l_j}
=
\int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k}
\mathbb{E}\, \sum_{(y_1,\dots,y_i)\in(\eta\cap A)^i_{\ne}} \sum_{(y_{i+1},\dots,y_{i+j})\in(\eta\setminus A)^j_{\ne}}
\\ \hfill
{\bf 1} \left( x_1,\dots,x_{m_1}\in\upsilon_\eta(y_1), \dots ,
x_{n-l_j+1},\dots,x_{n}\in\upsilon_\eta(y_{i+j}) \right)
\\ \hfill
dx_1\dots dx_n
\end{multline*}
and $B_{k,i,m_1,\ldots,m_i},B_{n-k,j,l_1,\ldots,l_j}$ are the same combinatorial coefficients as in the proof of Theorem~\ref{thm:asympt_volApprox}.
Fix some $i,j$, and $m_1,\dots,m_i,l_1,\dots,l_j$. Using the Slivnyak-Mecke formula
\eqref{2352} twice we get
\begin{multline*}
\beta_{i,j,m_1,\ldots,m_i,l_1,\ldots,l_j}
=
\lambda^{i+j} \int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k} \int\limits_{(\mathbb{R}^d\setminus A)^{j}} \int\limits_{A^i}
\\ \hfill
\P \left( x_1,\dots,x_{m_1}\in\upsilon_{\eta\cup \tilde{y}_{i+j} }(y_1),\dots,
x_{n-l_j+1},\dots,x_{n}\in\upsilon_{\eta\cup \tilde{y}_{i+j}}(y_{i+j}) \right)
\\ \hfill
dy_1\dots dy_{i+j} \, dx_1\dots dx_n ,
\end{multline*}
where $\tilde{y}_{i+j}=\{ y_1,\ldots, y_{i+j} \}$.
By Fubini and Lemma~\ref{2301},
\begin{multline*
\beta_{i,j,m_1,\ldots,m_i,l_1,\ldots,l_j}
\leqslant
\lambda^{i+j} \int\limits_{(\mathbb{R}^d\setminus A)^{j}} \int\limits_{A^i}
\\ \hfill
\prod_{r=1}^i\int\limits_{(\mathbb{R}^d\setminus A)^{m_r}}\left( \P\left( x_1,\dots,x_{m_r}\in\upsilon_{\eta\cup\{y_r\}}(y_r)
\right)\right)^{1/(i+j)} \, dx_1\dots dx_{m_r}
\\ \hfill \times\prod_{r=1}^j\int\limits_{A^{l_r}}\left( \P\left( x_1,\dots,x_{l_r}\in\upsilon_{\eta\cup\{y_{i+r}\}}(y_{i+r}) \right)\right)^{1/(i+j)}
\, dx_1\dots dx_{l_r}
\\ \hfill
dy_1\dots dy_{i+j} .
\end{multline*}
Using Lemma~\ref{1534} and \eqref{0629}, we get asymptotically as $\lambda\to\infty$
\begin{multline*}
\beta_{i,j,m_1,\ldots,m_i,l_1,\ldots,l_j}\leqslant
C\mathop{\rm Per}\nolimits(A)^{i+j}
\lambda^{i+j+\sum_{r=1}^i(-m_r-1/d)+\sum_{r=1}^j(-l_r-1/d)} \\
=C\mathop{\rm Per}\nolimits(A)^{i+j}\lambda^{-n+i+j-(i+j)/d}.
\end{multline*}
The maximum order of $\lambda$ is achieved for $i=k$, $j=n-k$, and the next term of maximum order is achieved for $i+j=n-1$, which together with \eqref{0736} and \eqref{0942} implies
\begin{multline*}
u_k
\leqslant
\lambda^{n} \int\limits_{A^{n-k}}\int\limits_{(\mathbb{R}^d\setminus A)^k} \int\limits_{(\mathbb{R}^d\setminus A)^{n-k}} \int\limits_{A^k}
\P\left( x_r\in\upsilon_{\eta\cup \tilde{y}_{n}}(y_r),\; r=1,\dots,n\right) \, dy_1\dots dy_n \, dx_1\dots dx_n
\\ + C\mathop{\rm Per}\nolimits(A)^{n-1}\lambda^{-1-(n-1)/d}
\end{multline*}
asymptotically as $\lambda\to\infty$. It is clear that
$$
\P\left( x_r\in\upsilon_{\eta\cup \tilde{y}_{n}}(y_r),\; r=1,\dots,n\right)\leqslant\P\left(B^{x_r}_{\|x_r-y_r\|}\cap\eta=\emptyset,r=1,\dots,n\right).
$$
Therefore, by Lemma~\ref{1952}, asymptotically as $\lambda\to\infty$,
\begin{equation}\label{2139}
u_k\leqslant v_k+2\sum_{s<t}v_{k,s,t}+C\mathop{\rm Per}\nolimits(A)^{n-1}\lambda^{-1-(n-1)/d},
\end{equation}
where
\begin{multline*}
v_k
=
\lambda^{n} \int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k} \int\limits_{(\mathbb{R}^d\setminus A)^{n-k}} \int\limits_{A^k}
\exp\left(-\lambda\kappa_d\sum_{r=1}^n\|x_r-y_r\|^d\right)
\, dy_{1}\dots dy_{n} \, dx_1 \dots dx_n ,
\end{multline*}
and
\begin{multline*}
v_{k,s,t}
=
\lambda^{n} \int\limits_{A^{n-k}} \int\limits_{(\mathbb{R}^d\setminus A)^k} \int\limits_{(\mathbb{R}^d\setminus A)^{n-k}} \int\limits_{A^k}
\exp\left(-\frac{\lambda\kappa_d}{n+1}\sum_{r=1}^n\|x_r-y_r\|^d\right)
\\ \times \exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right)
\, dy_{1}\dots dy_{n} \, dx_1 \dots dx_n .
\end{multline*}
By the identity~\eqref{1411},
\begin{equation}\label{2138}
v_k=2^{-n}(\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta))^n.
\end{equation}
Let us estimate $v_{k,s,t}$. For instance, we assume that
$s\leqslant k$ and $t\geqslant k+1$ (other cases are treated in
the same way). In the same way as in the proof of Theorem~\ref{thm:asympt_volApprox}, we obtain by inequality~\eqref{0629}
\begin{eqnarray*}
v_{k,s,t}
&\leqslant &
C \mathop{\rm Per}\nolimits(A)^{n-2}\lambda^{2-(n-2)/d}
\int\limits_{\mathbb{R}^d} \int\limits_{\mathbb{R}^d\setminus A} \int\limits_{\mathbb{R}^d} \int\limits_{A}
\exp\left(-\frac{\lambda\kappa_d}{(n+1)}\left(\|x_s-y_s\|^d\right)\right)
\\ && \hskip2cm \times
\exp\left(-\frac{\lambda\kappa_d}{2^{2d+1}(n+1)}\left(\|x_s-y_t\|^d+\|x_t-y_s\|^d\right)\right)
\, dy_s\, dy_t \, dx_s \, dx_t
\end{eqnarray*}
as $\lambda \to \infty$. Furthermore, by~\eqref{0609},
$$
v_{k,s,t}\leqslant C\mathop{\rm Per}\nolimits(A)^{n-2}\lambda^{-(n-2)/d}
\int\limits_{\mathbb{R}^d\setminus A} \int\limits_{A}
\exp\left(-\frac{\lambda\kappa_d}{(n+1)}\|x_s-y_s\|^d\right) \, dy_s dx_s,\quad \lambda\to\infty,
$$
and applying~\eqref{0629} again, we get
$$
v_{k,s,t}\leqslant C\mathop{\rm Per}\nolimits(A)^{n-1}\lambda^{-1-(n-1)/d},\quad \lambda\to\infty.
$$
Combining this with \eqref{2139} and \eqref{2138}, we
get
\begin{equation}\label{1249}
u_k\leqslant2^{-n}(\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta))^n+C\mathop{\rm Per}\nolimits(A)^{n-1}\lambda^{-1-(n-1)/d},\quad \lambda\to\infty.
\end{equation}
Inserting this into \eqref{0032} we otain
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A\Delta A_\eta)\leqslant (\mathbb{E}\mbox{\rm Vol}(A \Delta
A_\eta))^n+C\mathop{\rm Per}\nolimits(A)^{n-1}\lambda^{-1-(n-1)/d},\quad \lambda\to\infty.
$$
The application of Lyapunov's inequality
$$
\mathbb{E}\, \mbox{\rm Vol}^n(A\Delta A_\eta)\geqslant \left( \mathbb{E}\mbox{\rm Vol}(A\Delta
A_\eta)\right)^n
$$
finishes the proof.
\end{proof}
\begin{proof}[Proof of Corollary~\ref{cor:var}]
As was mentioned above, the second inequality immediately follows from Theorem~\ref{0037}. To prove the first one, let us again combine~\eqref{1249} and \eqref{0032} now for $n=2$ and $k=0,1,2$. We get for sufficiently large $\lambda$
$$
\mathbb{E}\,\mbox{\rm Vol}^2(A\setminus A_\eta)+\mathbb{E}\,\mbox{\rm Vol}^2(A_\eta\setminus A)\leqslant\frac12(\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta))^2+2C\mathop{\rm Per}\nolimits(A)\lambda^{-1-1/d},
$$
$$
2\mathbb{E}\,\left( \mbox{\rm Vol}(A\setminus A_\eta)\mbox{\rm Vol}(A_\eta\setminus A)\right)\leqslant\frac12(\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta))^2+2C\mathop{\rm Per}\nolimits(A)\lambda^{-1-1/d}.
$$
Combining this with Lyapunov's inequality
$$
\mathbb{E}\,\mbox{\rm Vol}^2(A\setminus A_\eta)+\mathbb{E}\,\mbox{\rm Vol}^2(A_\eta\setminus A)+2\mathbb{E}\,\left( \mbox{\rm Vol}(A\setminus A_\eta)\mbox{\rm Vol}(A_\eta\setminus A)\right)\\\geqslant(\mathbb{E}\mbox{\rm Vol}(A \Delta A_\eta))^2,
$$
we obtain for sufficiently large $\lambda$
$$
\mathbb{E}\,\mbox{\rm Vol}^2(A\setminus A_\eta)+\mathbb{E}\,\mbox{\rm Vol}^2(A_\eta\setminus A)-2\mathbb{E}\,\left( \mbox{\rm Vol}(A\setminus A_\eta)\mbox{\rm Vol}(A_\eta\setminus A)\right)\\\leqslant4C\mathop{\rm Per}\nolimits(A)\lambda^{-1-1/d},
$$
which together with~\eqref{1543} completes the proof.
\end{proof}
|
\section{Observations}
\noindent
Dust enshrouded activity of a galaxy can be studied ideally by
mid--infrared (MIR) observations. To explore the origin of the nuclear
MIR emission of galaxies as being due to either active galactic nuclei
(AGN) or star formation, observations of high spatial resolution are
required.
\noindent
The nuclear MIR surface brightness is introduced as a quantitative
measurement for AGN and starburst activity. However, one is unable to
distinguish between both activity types using the nuclear MIR surface
brightness derived from 4m class telescopes, even when adopting the
theoretical diffraction limit of $0.7''$ (FWHM) of such telescopes
(cmp. small gray symbols in Fig.~\ref{surf.ps}). Since the PSF width
is twice as large as for a 8m class telescope and the point source
sensitivity is a factor 16 lower, it becomes more difficult for a 4m
to resolve starburst and the surface brightness of unresolved sources
is reduced. Data recently obtained at 8m class telescopes
(Siebenmorgen et al. 2008) show that, out to a distance of 100Mpc, the
MIR surface brightness acquired clearly differentiate AGN from SB
behavior (Fig.~\ref{surf.ps}). Utilizing VISIR at the VLT the AGN
still appear point like whereas most starburst are resolved in the
MIR. This discrimination was made possible by an increase in spatial
resolution by a factor 2. Therefore it provides a clue to what will be
possible by increasing the spatial resolution by another factor 5 when
going from the VLT to the proposed extreme large telescope such as the
E-ELT which will be 40m class. For the E-ELT a mid infrared instrument
is included in the instrumentation plan and it has beside imaging also
high resolution spectroscopic and polarimetric observing capabilities
(Brandl et al., 2010).
\begin{figure}[htb]
\begin{center}
\includegraphics[width=11cm,angle=0]{extin.ps}
\caption{Mean extinction curve of the ISM by Fritzpatrick\&Massa
(2007, black) and a fit (magenta) by the dust model of
Sect.~\ref{dust}. Individual dust components are shown as
labeled. The grey areas indicate the 1$\sigma$ deviation as of the
samples.
\label{dust.ps} }
\end{center}
\end{figure}
\section{Dust model \label{dust}}
\noindent
Teams interested in modelling the processing of radiation by dust in
galaxies often apply a dust model as derived for the diffuse ISM of
the Milky Way. Dust cross sections are computed using similar optical
constants and temperature fluctuating particles such as PAHs are
included. We developed one of such dust models in which {\it {large}}
($60\rm{\AA}<a<0.2-0.3\mu$m) silicate (Draine 2003) and amorphous
carbon (Zubko et al. 2004) grains and {\it {small}} graphite
($5\rm{\AA}<a<80$\,\AA) grains are considered. We apply a power law
size distribution: $n(a) \propto a^{-3.5}$ and absorption and
scattering cross-sections are computed with Mie theory. In addition
there are {\it {PAHs}} with 30 and 200 C atoms with absorption cross
section as given by Schutte et al. (1993). By computing cross
sections above 100\,eV, we consider an approximation of kinetic energy
losses (Dwek\& Smith 1996) and apply it to all particles. The choice
of parameters is set up to achieve a fit of the mean extinction curve
of the ISM (Fitzpatrick \& Massa 2007), as shown in Fig.\ref{dust.ps}.
\noindent
With the advent of {\it{ISO}} and {\it{Spitzer}} more PAH emission
features and more details of their band structures are detected
(Tielens 2008). We consider 17 emission bands and take Lorentzian
profiles (Siebenmorgen et al. 1998). Parameters of what we call
astronomical PAH are calibrated using mid-IR spectra of starburst
nuclei and the RT model as of Sect.~{\ref{sb}}. PAH cross sections of
the emission bands are listed by Siebenmorgen \& Kr\"ugel (2001). In
the model, we use dust abundances of [X]/[H] (ppm) of: 31 for [Si],
150 [amorphous C], 50 [graphite] and 30 [PAH], respectively; which is
in agreement with cosmic abundance constraints (Asplund et
al. 2009). We are in the process of upgrading the model to be
consistent with the polarization of the ISM (Voshchinnikov
2004). Besides extinction the model accounts for the diffuse emission
of solar neighborhood when the dust is heated by the interstellar
radiation field (Mathis et al. 1983).
\section{Starbursts \label{sb}}
\noindent
There are three different ways in the literature trying to reproduce
the SED of extra-galactic nuclei. A first one uses a SED of a
well-known galaxy as a template to match other objects (Lutz et
al. 2002). A second group reproduce the shape of the SED by optical
thin dust emission using a scaled up version of the interstellar
radiation field (Draine \& Lee 2007). A third and more ambitious
method is to solve the radiative transfer (RT) problem using
assumptions about the galaxy. The latter is done at various levels of
sophistication and it may be instructive to point out technical
differences. Teams solving the RT problem evaluate the emission from a
dusty medium of spheroidal shape filled with stars and dust. At first
glance, the model results appear to agree, but upon closer inspection
one finds that deviations of derived parameters are substantial. For
example for Arp220 Grooves et al. finds an optical depth of a few
whereas we derive values between 70 -- 120. We admit that we did not
always find it easy to pin down exactly which approximations our
colleagues used, still we try to summarize the main features of some
RT models which are in widespread use. Monte Carlo techniques are
discussed in Sect.~\ref{agn}. We use the term {\it {dust
self--absorption}} when photons which are emitted by a dust
particles may be absorbed by other dust particles within the model
sphere and we call {\it {exact RT}} when the RT equation is solved
accurately including multiple scattering and dust self--absorption.
{$-$} Groves (2004, this volume): Shock, photoionization and dust
radiative transfer code called MAPPINGIII. The dust is distributed in
a screen and the RT is solved in one dimension ignoring dust
self-absorption and treating scattering in forward direction only.
{$-$} Efstathiou \& Rowan--Robinson (2003), Efstathiou (this volume):
Dust is distributed in spherical symmetry. There are two components:
molecular clouds with exact RT computation and cirrus which is added
as a foreground screen. Both components are uncoupled in the RT
equation. The code includes a free parameter to scope with the
observed optical and UV spectrum.
{$-$} Takagi et al. (2003, this volume): Exact RT in spherical
symmetry where dust and stars are homogeneously distributed. Molecular
clouds (clumps) are not treated. The code includes a free parameter to
scope with the observed optical and UV spectrum.
{$-$} Siebenmorgen et al. (2007): Exact RT in spherical symmetry where
dust and a young and old population of stars are distributed in the
galaxy. A fraction of the young stars are in molecular clouds for
which a second exact RT computation is solved. The coupling of both RT
solutions, that of the galaxy and the embedded sources, is treated and
this is a particular feature of the model.
{$-$} Silva et al. (1998): Present a code called GRASIL in which dust
is distributed in axial-symmetry. There is a molecular cloud
component with exact RT. A cirrus component is added in which the RT
is solved by ignoring dust self-absorption and in which dust
scattering is simplified by altering the dust absorption cross
section. In addition the code includes a free parameter to scope with
the observed optical and UV spectrum. All three components are
uncoupled in the RT equation.
{$-$} Popescu et al. (2002, this volume): Dust is distributed in
axial-symmetry. The model is fine tuned to Spiral galaxies
(Sect.~\ref{spirals}) and includes a bulge, two galactic disk
components and molecular clouds (clumps). The RT is solved by
ignoring dust self-absorption and only the first scattering event is
treated. Clumps are added by a pre-computed template spectrum. There
is a free parameter to scope with the observed optical and UV
spectrum. All five components are uncoupled in the RT equation.
\noindent
As the coupling between the RT calculations of the galaxy and the
embedded sources is a computer intensive problem we provide a library
of some 7000 SEDs {\footnote{SED library available at: {\tt
{http://www.eso.org/$\sim$rsiebenm/sb\_models/}}}} for the
nuclei of starburst and ultra--luminous galaxies (Siebenmorgen et al.,
2007). Its purpose is to quickly obtain estimates of the basic
parameters of the object, such as luminosity, size and dust or gas
mass and to predict the flux at yet unobserved wavelengths using a
physical model. Unfortunately, for faint or red-shifted objects
photometry is sometimes only provided at two MIR bands, for example at
8 and 24$\mu$m from Spitzer. In Fig.~\ref{2dat.ps} we demonstrate for
the ULIRG NGC6240 that the SED library can be used to estimate the
total luminosity to within a factor $\sim 2$ in case only two such MIR
fluxes are known. We also see (Fig.~\ref{2dat.ps}) that the SED will
be quite well constrained by an additional submm data point. The
model is applied to high red shifts ($z \approx 3$, Efstathiou \&
Siebenmorgen 2009) where PAH have been detected.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=8cm,angle=90]{2datN6240.ps}
\caption{Elements of the SED starburst library by Siebenmorgen et
al. (2007, dotted) which fit photometry (circles) of NGC6240 at
8$\mu$m (Siebenmorgen et al. 2004) and 24$\mu$m (Klaas et al. 2001)
to within 30\%; submm data by Benford (1999) and Klaas et
al. (2001), best fit is indicated as thick line \label{2dat.ps}. }
\end{center}
\end{figure}
\section{Spirals \label{spirals}}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=10.5cm,angle=270]{scheSpiral.ps}
\end{center}
\vspace*{-1.0 cm}
\caption{Schematic view of the geometry applied in the models
of Spirals by Popescu et al. \label{scheSpiral.ps} }
\end{figure}
\noindent
The geometrical distribution of dust in spiral galaxies has been
investigated by fitting images in the optical/NIR. This is done by
means of 2d radiative transfer codes which include a bulge component
and an old stellar population of stars; originally only dust
absorption and scattering was treated (Kylafis \& Bahcall 1987,
Xilouris et al. 1999, Misiriotis et al., 2000). Building into the
models typical dust emission properties gives a puzzle when compared
to observations in the FIR. Typically the models underestimate the FIR
luminosities by a factor of 3. The FIR excess could be explained by
Popescu et al. (2000, 2010, this volume) introducing more components.
In their models there is a distribution of diffuse dust associated
with the old and young stellar disk populations as well as a clumpy
component arising from dust in the parent molecular clouds in star
forming regions. Basic parameters of their models are: angular size
and inclination of the disk, the central face-on dust opacity in the
B-band, a clumpiness factor for the star-forming regions, the
star-formation rate, the normalized luminosity of the stellar
population and the bulge-to-disk ratio and a wavelength dependent
escape probability of stellar radiation (which is sometimes treated as
model output). A schematic view of the geometry of the RT model is
presented in Fig.~\ref{scheSpiral.ps}. The model is successfully
applied to a large sample of galaxies from the Millennium Galaxy
Catalog Survey (Driver et al. 2007). The observed
attenuation--inclination relation is fit when a two dust disks are
considered whereas data are not fit in a single disk scenario.
Over the past decade the edge on spiral NGC891 become a benchmark test
for RT models of spiral galaxies. Bianchi (2008) uses a clumpy disks
model and applied MC techniques which are of advantage when dealing
with such complicated geometries (see Sect.~\ref{agn}). His results
give support to the idea that the diffuse dust disk is more extended
than the stellar disks. De Looze et al. (20011, this volume) present a
MC radiative transfer model of the Sombrero galaxy which is able to
fit the SED and optical/NIR images and extinction profiles. Using only
an old stellar population the dust luminosity in the FIR is again
underestimated by a factor 3. The discrepancy is solved by assuming a
star forming stellar component both in the ring and inner disk to
account for emission at 24 and 70$\mu$m. In the submm an additional
dust component is used, accounting for 75\% of the total dust content
and distributed in quiescent compact clumps. A possibility that part
of the dust could be composed of grains with a higher submm
emissivity, of for example large fluffy grains (Kr\"ugel \&
Siebenmorgen 1994), could not be ruled out. On the other hand a
similar approach introducing a clumpy medium in a single disk model is
applied to the edge-on spiral galaxy UGC4754 (Baes et al. 2011). The
model has a deficit in the FIR luminosity and the authors propose
higher submm dust emissivities to solve the energy ballance at such
long wavelengths.
\section{AGN and Monte Carlo \label{agn}}
\noindent
Dust is detected in the majority of active galactic nuclei (Haas et
al., 2008). According to the unified scheme (Antonucci \& Miller,
1985), AGN are surrounded by a dust obscuring torus. However,
observations are not able to resolve the inner parts of AGNs so that
the geometrical distribution of the dust is a matter of
debate. Theoretical considerations favor a clumpy structure in a torus
like configuration of optical thick dust clouds surrounding the black
hole accretion disk (Pier \& Krolik (1992). Some evidence of a clumpy
or filamentary structure of the torus is given by VLTI observations of
the nearby active galactic nuclei in Circinus (Tristram et al. 2007).
Radiative transfer models of homogeneous toroidal structures
over-predict the silicate absorption and emission band at around
10$\mu$m when compared to observations. The silicate emission feature
in type I AGN is rather shallow (Siebenmorgen et al. 2005). A
statistical attempt to describe the radiation from a clumpy AGN torus
is given by Nenkova et al. (2002) and more detailed radiative transfer
computations using the Monte Carlo technique are presented by H\"onig
et al. (2006) and Schartmann et al. (2008).
We develop a vectorized three dimensional Monte Carlo (MC) technique
to solve the radiative transfer problem in such a fairly complicated
geometry of a clumpy dust torus. Originally the MC radiative transfer
method for scattered light in astronomical sources is presented by
Witt (1977) and important improvements are introduced by Lucy
(1999). A first version of the code which we are using is developed by
Kr\"ugel (2006). To reduce the computational effort of MC dust
radiative transfer methods different optimization strategies are
developed (Lucy 1999, Bjorkman \& Wood 2001, Gordon et al. 2001,
Misselt et al. 2001, Baes 2008, Bianchi 2008, Baes et al. 2011). A
numerical solution of the radiative transfer equation which is
specifically developed to be vectorized and to run on graphical
computer units (GPU) is presented by Heymann (2010). A GPU version
including stochastic heated grains is presented (Siebenmorgen et
al. 2010), in which some detailed physics of the destruction of PAHs
by soft (photospheric) and hard (X--ray) radiation components is
treated, (see Siebenmorgen \& Kr\"ugel 2010 for a discussion of PAH
dissociation). The speed up factor of the GPU method is proportional
to the number of processors available. Therefore the GPU scheme is on
our standard PC with a conventional graphic card about 100 times
faster when compared to the original scalar version of the MC program.
The numerical solution provides the temperature of the dust, spectra
and images within acceptable timescales. The code is tested against
existing benchmark results. The method handles arbitrary dust
distributions in a three dimensional Cartesian model space at various
optical depths. Therefore it is well adopted to be applied to a clumpy
torus structure around an AGN.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=6cm,angle=0]{agnTklump.eps}
\end{center}
\caption{Zoom into the clumpy AGN torus displaying the temperature
structure of individual clouds. \label{agnTklump} }
\end{figure}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=13.5cm,angle=0]{torus_emm_sca.eps}
\caption{Intensity maps of the clumpy AGN dust torus model
(color-code: cgs units normalized to the peak intensity). The model
is made up of 1000 clumps and each clump has a total optical depth
of $\tau_{\rm V} = 30$. Mid IR images at 10\,$\mu$m are dominated
by dust emission and displayed in the top panel. The optical light
at 0.55\,$\mu$m represents the scattering intensities and are shown
in the bottom panel. The AGN torus is viewed face-on ($\theta =
0^\circ$, left), at $\theta = 45^\circ$ (middle) and edge-on
($\theta = 90^\circ$, right). \label{agn.ps} }
\end{center}
\end{figure}
The primary heating source of the dust in the torus emerges from the
accretion disc around the massive black hole. For the spectral shape
we apply a broken power law as suggested by Rowan-Robinson (1995). As
example we use a total AGN luminosity of $L_{\rm {AGN}} =
10^{45}$\,erg/s. This sets the inner radius $R_{\rm {in}} \sim 0.4
\sqrt{L_{\rm {AGN}}/10^{45}}$\,[pc] where the dust evaporation zone is
located. We treat the torus to an outer radius of $R_{\rm {out}} = 50
R_{\rm {in}}$ and consider a torus opening angle of $θ \sim 20^{\rm
o}$. Clouds are randomly distributed in the otherwise optical thin
dust torus. Individual clouds are approximated with a sphere of
constant density and radius of 3\,pc. The optical depth through the
center of the clouds is $\tau_{\rm V} = 30$. In Fig.~{\ref{agnTklump}}
we zoom into a part of the clumpy torus and show the temperature
structure of the clouds. The individual hemispheres which are oriented
in direction of the central heating source are warmer as compared to
their dark sides. Shadowed clumps become also colder than those
directly heated by the central engine. Intensity maps of the emission
at 10$\mu$m and scattered light at 0.5$\mu$m of the clumpy AGN dust
torus model are displayed in Fig.~\ref{agn.ps}. The maps show that in
a clumpy medium even in the edge on case some radiation from the
central region is visible. A more detailed description of the clumpy
AGN torus model and a further application to the silicate emission in
Quasars is given in Heymann \& Siebenmorgen (2011).
|
\section{Introduction}
\label{sec:intro}
During the last decade there has been an increasing observational effort to understand the nature of the HCN emission from galaxies.
This includes studies of HCN emission in nearby AGN and starbursts \citep{koh01,koh05,koh08,kri07,kri08} as well as luminous and ultraluminous galaxies \citep{gra06,gra08}. One result is that HCN line emission is enhanced with respect to CO in at least some AGN compared with starbursts \citep{koh05,koh08}. The high HCN/CO intensity ratios could be due to high molecular gas densities near the AGN, and/or high
HCN/CO abundance ratios that might be due to elevated X-ray ionisation and heating rates near the accreting black holes.
For example, in an early study, \cite{ste94} concluded that the large HCN/CO intensity ratio observed in the nuclear ($< 100$ pc) region
in the Seyfert-2 galaxy NGC 1068 indicates a high HCN/CO abundance ratio $\sim 10^{-2}$ near the AGN.
\cite{lep96} and \cite{use04} argued that elevated HCN/CO abundance ratios are signatures of high X-ray ionisation rates \citep{mal96,mei05,bog05,mei07}. High gas densities
also likely play a role in boosting the HCN line intensity \citep{kri08}.
In this paper we present high resolution observations of CO (2-1) and HCN (1-0) rotational line emissions in the inner regions of
NGC\,3227, a nearby (D=17\,Mpc; 1\arcsec $\sim$ 80\,pc) Seyfert galaxy for which detailed studies have been made of the stellar \citep{dav06,dav07} and gaseous \citep{sch00,hic09} content of its central regions. We analyze the data using
large-velocity-gradient (LVG) computations.
For this galaxy we again find a significant enhancement in the HCN/CO intensity ratio close to the active nucleus.
The observed nuclear intensity ratio is consistent with optically thick thermalized emission in dense ($\gtrsim 10^5$~cm$^{-3}$) gas.
Alternatively, and especially in comparison with similar data in NGC 1068 and NGC 6951, the nuclear emissions could be
tracing optically thinner emission in which the HCN/CO abundance ratio is large.
\section{Observations}
\label{sec:data}
The analysis in this paper is based on subarcsecond resolution observations of the CO(2-1) 230.5\,GHz and HCN(1-0) 88.6\,GHz lines.
The CO(2-1) data, for which the beam is 0.6\arcsec, were previously presented by \cite{sch00}.
New 3\,mm HCN(1-0) data, with a 0.9\arcsec$\times$1.2\arcsec\ beam, were obtained during February 2009 in the A configuration (760\,m baseline) of the six 15-meter antennas of the IRAM Plateau de Bure Interferometer.
The H$^{12}$CN(1-0) line at 88.6\,GHz and the H$^{13}$CN(1-0) line at 86.3\,GHz were observed together, using a single polarisation for each 1\,GHz bandwidth segment.
The system temperature was 80--100\,K.
Atmospheric conditions were moderate, with winds and $\sim5$\,mm precipitable water vapour.
Phase and amplitude variations were calibrated out by interleaving
reference observations of standard calibration sources.
The data were processed and calibrated using the CLIC program in the
IRAM GILDAS package, and binned spectrally to a resolution of
50\,km$^{-1}$ to increase the signal to noise.
The kinematics of the HCN(1-0) line emission in NGC\,3227 are analysed by \cite{san11} together with several other galaxies.
It is remarkable that despite the similar beam sizes, the CO(2-1) map clearly shows the circumnuclear ring at a radius of 140\,pc (1.7\arcsec), while the HCN map shows emission from only the central region, within a few tens of parsecs from the nucleus.
This can be seen clearly in Fig.~\ref{fig:obs} and is reflected in the line fluxes given in Table.~\ref{tab:data}.
In the nuclear region, which we define here to be the central 1\arcsec\ (i.e. $r<40$\,pc), the flux ratio is F$_{HCN1-0}$/F$_{CO2-1}$=0.11;
on the other hand in the circumnuclear ring, for which we take an annulus 2--4\arcsec, we find F$_{HCN1-0}$/F$_{CO2-1}$=0.01, an order of magnitude less.
Since the emission is spatially resolved in both cases, this difference cannot be due to differing beam dilution effects.
Instead it is likely to be due to differences in either molecular abundance -- e.g., as a result of a distance dependent X-ray ionisation rate -- or differences in excitation efficiencies -- e.g., due to higher nuclear gas densities favoring HCN. (The critical H$_2$ density of $\sim 7\times 10^{3}$~cm$^{-3}$
for collisional deexcitation of CO (2-1)
is significantly lower that the $\sim 3\times 10^6$~cm$^{-3}$ for HCN (1-0).)
In the next section we use large velocity gradient (LVG) models to address these issues.
\begin{figure}
\centering
\includegraphics[angle=0,width=8.8cm]{plot2new.eps}
\caption{\label{fig:obs}
Images of the CO(2-1) line emission \citep[from][]{sch00} and HCN(1-0) line emission from NGC\,3227.
Overdrawn is the ellipse tracing the circumnuclear ring \citep{dav06}.
While the CO emission originates mostly from the ring, the HCN intensity is far higher in the central region. There is an order-of-magnitude change in the flux ratio between these regions.
}
\end{figure}
\begin{table}
\caption{CO and HCN fluxes in NGC3227}
\label{tab:data}
\begin{tabular}{l c c c}
\hline\hline
aperture & F$_{HCN1-0}$ & F$_{CO2-1}$\tablefootmark{a} & F$_{HCN1-0}$/F$_{CO2-1}$\\
& Jy\,km\,s$^{-1}$ & Jy\,km\,s$^{-1}$ & \\
\hline
1\arcsec & $0.48\pm0.06$ & $\phantom{0}4.4\pm\phantom{0}1.3$ & $0.114^{+0.051}_{-0.035}$ \\
2\arcsec & $1.37\pm0.17$ & $19.5\pm\phantom{0}5.9$ & $0.074^{+0.033}_{-0.023}$ \\
4\arcsec & $1.86\pm0.29$ & $57.6\pm17.3$ & $0.033^{+0.016}_{-0.011}$ \\
2--4\arcsec & $0.49\pm0.33$ & $38.5\pm18.3$ & $0.013^{+0.024}_{-0.008}$ \\
\hline
\end{tabular}
\tablefoottext{a}{Scaled from the 75\,mJy given by \cite{sch00} in a 8.4\arcsec$\times$8.4\arcsec\ aperture and with an uncertainty of 30\%.}
\end{table}
\section{Physical Properties of the Molecular Gas}
\label{sec:gasprop}
\subsection{LVG Calculations}
\label{sec:lvg}
\begin{figure*}
\centering
\includegraphics[angle=0,width=\textwidth]{hcn10co21_recalc.eps}
\caption{\label{fig:lvg}
LVG calculations for a 4-dimensional parameter space: the [HCN]/[CO] abundance ratio and gas kinetic temperature are given at the top of each panel, while the axes of each panel are the gas volume density and ratio between the column density and linewidth (or equivalently between the volume density and velocity gradient). The contours show the expected line ratio based on the emitted fluxes in Jy\,km\,s$^{-1}$.
The locus of representing possible parameters for the nucleus and ring of NGC\,3227 are drawn in red and green respectively.
The dashed orange lines are curves of equal velocity gradient (in km\,s$^{-1}$\,pc$^{-1}$), indicating which regions of the parameter space are physically plausible.
In particular, the locus for self-gravitating virialised clouds is represented by the dot-dash blue line.
}
\end{figure*}
We have constructed a new LVG code
and have used it to compute the HCN (1-0) and CO (2-1) line intensities for a wide range of conditions.
We calculate the line source functions assuming photon escape probabilities from spherical clouds. We use the recent \cite{yan10} data for
the excitations and deexcitations of the CO rotational levels that are induced by collisions with H$_2$.
For HCN we use the \cite{gre74} collisional data, as updated and listed in
the RADEX database \citep{tak07}. We have cross-checked all of our results with the LVG code RADEX,
and find excellent agreement:
in the parameter space we have assessed, the converged line ratios agree to within 1.5\%.
We assume that the HCN and CO molecules are mixed uniformly, and that the corresponding line emissions arise from gas at the same temperature and density.
The luminosity ratios resulting from the model calculations are presented graphically in Fig.~\ref{fig:lvg}, covering the parameter space:
kinetic temperature $30 \leq T\,[K] \leq 300$,
HCN to CO abundance ratio $10^{-5} \leq X_{HCN}/X_{CO} \leq 10^{-2}$
(with $X_{CO}=10^{-4}$ is the CO abundance relative to hydrogen),
H$_2$ volume gas density $10^{3} \leq n_{H_2}\,[cm^{-3}] \leq 10^{7}$,
and a ratio of gas-density to velocity-gradient, or equivalently
column density to linewidth, of
$5\times10^{17} \leq N_{H_2}/dV\, [cm^{-2}\,(km\,s^{-1})^{-1}] \leq 5\times10^{23}$.
In each panel in Fig.~\ref{fig:lvg}, the parameter-space consists of four regimes.
The upper right parts
correspond to local thermal equilibrium (LTE) in the optically thick limit.
In this regime
the transition excitation temperatures $T_{ex}$ of the lines approaches the kinetic temperature $T_{kin}$ of the gas.
Since the line flux, in the Rayleigh-Jeans limit, is
$F \propto \int T_{ex} \nu^2 d\nu$, and an interval $dV$ in velocity space is
$dV = c \, d\nu / \nu$, the ratio of two line fluxes measured as
$\int F \, dV$ (e.g. in units of Jy\,km\,s$^{-1}$ as used here) is
$F_1/F_2 = (\nu_1/\nu_2)^2$
in the optically thick and LTE limit.
This corresponds to 0.15 for the ratio of the HCN(1-0) and CO(2-1) lines at 88.6\,GHz and 230.5\,GHz respectively,
which is close to the observed ratio of 0.1 in the nucleus.
The upper left regions also correspond to LTE because the gas densities are high, but
in this regime the line optical depths are low because the velocity gradients (or line widths) are large.
For such conditions, the line intensities are linearly proportional to the molecular abundances.
This behavior is reflected in the HCN/CO intensity ratios indicated by the contour values.
On the left sides of the panels, the intensity ratios decrease linearly with
the assumed abundance ratio
$X_{HCN}/X_{CO}$, which ranges from $10^{-2}$ to $10^{-5}$ in Fig.~\ref{fig:lvg}.
The lower parts of each panel correspond to low densities for which the HCN
is subthermally excited, leading to relatively higher populations in the lowest rotational levels.
Again the optical depth increases from left to right as the column density increases for fixed line width.
Some regions of the parameter space may be less physically plausible
because they assume very large velocity gradients.
Lines of constant velocity gradient are indicated by the dashed orange lines in Fig.~\ref{fig:lvg},
with increasing gradients towards the upper left.
For self-gravitating virialised clouds $\Delta V/R \sim n^{1/2}$ where $\Delta V$ is the velocity dispersion, $R$ is the cloud radius,
and $n$ is the gas density. Treating $\Delta V/R$ as a velocity gradient \citep[e.g.][]{gol01} gives
$dV/dr \sim {\rm 3.1\,km\,s^{-1}\,pc^{-1}} \sqrt{n_{H_2}/10^4\,{\rm cm^{-3}}}$, or
typically a few~km\,s$^{-1}$\,pc$^{-1}$, or a few tens in cases of extreme density.
In Fig.~\ref{fig:lvg}, the virial relation $n_{H_2} \propto (N_{H_2}/dV)^2$ is represented by the dot-dash blue line in each panel.
Clouds that are unbound or at least partially pressure confined could be to the left of this line.
\subsection{Analysis for NGC\,3227}
\label{sec:nuc}
Curves representing the possible parameter ranges for the measured flux ratios F$_{HCN1-0}$/F$_{CO2-1}$ for NGC\,3227 have been overdrawn for the nucleus (red lines) and circumnuclear ring (green lines), as defined in Sec.~\ref{sec:data}.
If we assume that the clouds are self-gravitating, we are restricted to the points where the red and green lines intersect the dashed blue line.
These show that if the clouds in the ring have a similar density to the nucleus, then the HCN abundance must be about 2 orders of magnitude lower.
Alternatively, the density may differ by up to 2 orders of magnitude if the abundances are similar.
We discuss this further in Sec.~\ref{sec:xray}
In the nucleus, the flux ratio F$_{HCN1-0}$/F$_{CO2-1}$=0.11 is remarkably close to the LTE optically thick limit, and its locus includes a contour around this region.
However, depending on the HCN abundance, the gas density at which this occurs can vary from $n_{H_2} \sim 10^4$\,cm$^{-3}$ at the highest abundance to
$\sim 3\times10^5$\,cm$^{-3}$ at the lowest abundance we have considered,
with column density
$N_{H_2}/dV \gtrsim 10^{22}$\,cm$^{-2}$/(km\,s$^{-1}$).
By considering the mean volume density of the HCN emitting region, and putting a limit on a realistic filling factor, we can restrict this range further.
The first step is to estimate the volume of the emitting region. This can be done because the HCN emission is marginally resolved.
A detailed estimate of the intrinsic size -- via dynamical modelling, and taking into account emission from the ring -- is given in \cite{san11}.
These authors show that the diameter is 0.54\arcsec, corresponding to 45\,pc, and the scale height is 6\,pc.
Taking this as an indication of the size along the line of sight sets the volume of the emitting region.
The second step is to estimate the mass.
We have done this in several ways since they are all uncertain.
\begin{enumerate}
\item
The LVG calculation yields a mass directly under the assumption that the ratio of the observed linewidth to the linewidth of an individual cloud is tracing the number of clouds, such that
$N_{tot} = N_{cloud} (\delta\nu_{obs} / \delta\nu_{cloud})$.
For conditions corresponding to
$X_{HCN}/X_{CO} = 10^{-2}$,
T$=300$\,K,
$n_{H_2} = 10^{5.5}$\,cm$^{-3}$ and
$N_{H_2}/dV = 10^{22}$\,cm$^{-2}$\,(km\,s$^{-1})^{-1}$ (see Sec.~\ref{sec:others}), the observed HCN(1-0) flux leads to a mass of
$1.8\times10^6$\,M$_\odot$.
\item
The standard method to estimate the mass is from the CO luminosity.
We have done this directly from the CO(2-1) line flux in Table~\ref{tab:data} using a conversion factor $\alpha = 4.3$ which includes a correction for helium \citep{tac08}.
This yields $3.3\times10^6$\,M$_\odot$.
\item
A similar conversion for the HCN line that has been calibrated by \cite{kri08} for AGN is
$M_{H_2}$/L$_{HCN} \sim 10\,M_\odot$\,(K\,km\,s$^{-1}$\,pc$^2$)$^{-1}$.
This yields a mass of $6\times10^6$\,M$_\odot$.
\item
As a final check, we use the dynamical mass derived from the HCN kinematics.
By fitting models to account for the beam smearing, \cite{san11} find $M_{dyn} = 5.6\times10^7$\,M$_\odot$.
For a nominal 10\% gas fraction expected in local disks and starbursts \citep{hic09}, this would yield a gas mass of $6\times10^6$\,M$_\odot$.
\end{enumerate}
These estimates are all the same order of magnitude, and suggest that the gas mass in the central arcsec is of order $4\times10^6$\,M$_\odot$.
Hence we can estimate the mean density to be
$\langle n_{H_2}\rangle \gtrsim 6 \times10^3$\,cm$^{-3}$.
Comparing this to the cloud densities above yields volume filling factors in the range 1--0.01.
In this range, a lower filling factor is more physically plausible,
which would tend to favour the solutions with higher cloud densities.
Fig.~\ref{fig:lvg} shows these have either higher temperature or less extreme HCN abundance.
\subsection{Comparison to NGC\,6951 and NGC\,1068}
\label{sec:others}
NGC\,1068 and NGC\,6951 are two other galaxies for which the HCN(1-0)/CO(2-1) ratio has been measured on comparable $\sim100$\,pc scales.
We use flux densities reported by \cite{kri07} for the nuclear region (denoted `C' in their Table~1) of NGC\,6951; and also the values for the circumnuclear disk of NGC\,1068, as the sum of the red and blue channels reported in Table~3 of \cite{use04}.
These yield line ratios (for line fluxes in Jy\,km\,s$^{-1}$) of $0.37\pm0.05$ and $0.214\pm0.002$ respectively, and are denoted by the solid magenta lines on Fig.~\ref{fig:lvg}.
These lines appear almost exclusively in the panels corresponding to the highest HCN abundance we have considered, $X_{HCN}/X_{CO}=10^{-2}$.
In contrast to NGC\,3227, in which the line emission appears to be optically thick, the loci of the magenta lines for NGC\,1068 and NGC\,6951 are toward the optically thin (left) side of the panels.
Despite this, it is notable that there are regions of the parameter space where the contours corresponding to all 3 objects lie close together, running from lower left to upper right.
The region extends from $n_{H_2}=10^{4}$\,cm$^{-3}$ and $N_{H_2}/dV=10^{19}$\,cm$^{-2}$/(km\,s$^{-1}$)
to $n_{H_2}=10^{6}$\,cm$^{-3}$.
It is precisely because one can attribute the observed line ratios -- with different optical depths for the 3 galaxies -- to similar physical properties of the gas in all these 3 objects that this locus is appealing.
\begin{figure}
\centering
\includegraphics[angle=0,width=9.0cm]{tau2.eps}
\caption{\label{fig:tau}
Optical depth $\tau$ of the HCN(1-0) line (left) and CO(2-1) line (right) for gas properties corresponding to the bottom left panel in Fig.~\ref{fig:lvg}.
Darker regions correspond to lower optical depth.
Overplotted are the contours for the 3 galaxies corresponding to their HCN(1-0)/CO(2-1) ratio.}
\end{figure}
Why this occurs can be seen in Fig.~\ref{fig:tau} which shows the optical depths $\tau$ for the HCN(1-0) and CO(2-1) transitions.
The gas properties of both these panels correspond to the bottom left panel in Fig.~\ref{fig:lvg} (300\,K and $X_{HCN}/X_{CO}=10^{-2}$), and cover the same range of density and velocity gradient.
These plots show clearly the characterisation of the different regions:
in the lower half the HCN(1-0) line is optically thick because the density is low enough that it is sub-thermal; above the critical density, the line is in LTE and thus optically thin at low columns and optically thick at high columns.
The locus where all the contours for the 3 galaxies are close together and parallel follows approximately the boundary where the HCN(1-0) line becomes optically thick.
Here, a small change in physical conditions (column or density) can result in the HCN(1-0) emission switching from optically thin to optically thick.
This regime is, however, also associated with very large velocity gradients.
It is $dV/dr\sim10^4$\,km\,s$^{-1}$\,pc$^{-1}$ at $T=30\,K$, but reduces as the temperature increases.
Velocity gradients were not discussed explicitly by \cite{ste94} or \cite{use04} in their $T=50$\,K LVG calculations for NGC\,1068.
But their analyses also associate the observed properties with similarly extreme velocity gradients.
Indeed, one of the main conclusions of \cite{ste94} was that
$X_{HCN}/X_{CO} \gtrsim 10^{-2}$ in NGC1068.
For the temperature they considered, this would lead to $dV/dr\sim10^4$\,km\,s$^{-1}$\,pc$^{-1}$ (matching the top left panel of Fig.~\ref{fig:lvg} here).
However, our LVG calculations shows that $dV/dr$ is reduced as both the temperature and density increase.
When considering all 3 galaxies together, the smallest -- and therefore arguably the most physically plausible -- value in the parameter space we have covered is
$dV/dr\sim100$\,km\,s$^{-1}$\,pc$^{-1}$ at
$T=300$\,K and $n_{H_2}\sim10^{5.5}$\,cm$^{-3}$.
This location is not far from the boundary of the optically thick LTE regime discussed previously, but due to the high velocity gradient represents clouds that are either pressure confined or unbound.
Interestingly, there is evidence in NGC\,1068 from recent Herschel observations with PACS of high rotational CO transitions, for a significant mass of molecular gas in the central $\sim100$\,pc at temperatures of 100\,K and 400\,K and densities of $\sim10^{6.5}$\,cm$^{-3}$ \citep{hai11}.
Similarly, in an analysis of various HCN, HCO$+$ and CO isotope transitions in the central 100\,pc of NGC\,1068, \cite{kri11} also argued in favour of warm (T$\gtrsim200$\,K) gas.
However, they also concluded that the density is of order $n_{H_2}\sim10^{4}$\,cm$^{-3}$.
Our LVG calculations indicate that such densities are associated with very high velocity gradients for the observed line ratio, which we consider physically unlikely.
This, combined with a comparison of the cloud density to mean density estimated in Sec.~\ref{sec:nuc} has led us to favour the higher density solution with more moderate velocity gradient.
\subsection{X-ray Ionisation rate in NGC\,3227}
\label{sec:xray}
Our LVG calculations presented above suggest that as one alternative the high HCN(1-0)/CO(2-1) ratio in the central
$\sim$100\,pc of NGC\,3227 could be due to an exceptionally high HCN abundance,
an interpretation supported by the even higher HCN/CO intensity ratios observed in the nuclei of
NGC\,1068 and NGC\,6951.
It is possible that the high HCN abundances are associated with elevated X-ray
ionisation and/or heating rates near the AGN.
Theoretical investigations of X-ray (or cosmic-ray) driven chemistry show that the
(steady-state) molecular abundances depend primarily on the ratio of the
cloud density $n$ to X-ray ionisation rate
$\zeta$ \citep{kro83,lep96,mal96,mei05,bog05,mei07}.
Here we adopt the notation of \cite{bog05},
normalising $\zeta$ to $10^{-17}$\,s$^{-1}$ to give the parameter
$n/\zeta_{-17}$. The density $n = n_H + n_{H_2}$ refers to the total
atomic plus molecular hydrogen density.
In the following analysis, we first calculate the ratio $n/\zeta_{-17}$ in the nucleus of NGC\,3227 and then compare it to ratios predicted by models of X-ray irradiated gas.
\subsubsection{$n/\zeta_{-17}$ in NGC\,3227}
To estimate the X-ray ionisation rate in the central $\sim$100\,pc of NGC 3227
we need to know the intrinsic spectral energy distribution from the AGN.
We adopt
the SXPL model of \cite{mar09} in which both the hard and soft
components of the X-ray flux are modelled with power laws:
\[
N_{\rm ph} = 0.0040(E/keV)^{-3.35} + 0.0067(E/keV)^{-1.57}
\]
where $E$ is the photon energy in keV and $N_{\rm ph}$ is the photon
flux in units of ph\,keV$^{-1}$\,cm$^{-2}$\,s$^{-1}$.
Making the usual assumption that the primary ionisation rate of
hydrogen is negligible compared to the secondary ionisation rate, we
then calculate the resulting ionisation rate $\zeta$ using equation
A4 of \cite{mal96}:
\[
\zeta = N_{\rm sec} \int^{E_{\rm max}}_{E_{\rm min}}
\sigma_{\rm pa}(E) \, F(E) \, dE.
\]
Here $N_{\rm sec}=28$
(as given by \citealt{mal96}) is the number of
secondary ionisations per keV of primary photoelectron energy
assuming a mean-energy per ion-pair of 37.1 eV for
energy deposition in a
molecular hydrogen gas (\citealt{dal99});
$\sigma_{\rm pa}(E)$ is the absorption cross section per
H nucleus, for which we adopt the broken power-law fit in equation A5
of \cite{mal96};
and $F(E)$ is the incident flux in units of ph\,keV$^{-1}$.
The photoionisation is dominated by
photons with energies for which $\tau\sim 1$.
We therefore take
the limits of the integral to be $E_{\rm max}=100$\,keV and
$E_{\rm min}$ as the energy at which the optical depth due to
photoelectric absorption is $\tau(E)=1$, ignoring attenuation above
this limit.
The energy at which $\tau(E)=1$ is interpolated from Table~9.3 of
\cite{sew00} for the \cite{mor83} model, for a given column density.
And the column density is assumed to be proportional to distance from
the AGN up to a maximum of $3\times10^{23}$\,cm$^{-2}$ (\citealt{hic09})
at 30\,pc.
We have evaluated the integral at two distances: 18\,pc, an area
weighted mean distance from the AGN to the HCN emitting gas that
corresponds to the nuclear region;
and 140\,pc corresponding to the distance of the circumnuclear ring.
We find
$\zeta_{\rm 18pc} = 3.6\times10^{-13}$\,s$^{-1}$ and
$\zeta_{\rm 140pc} = 3.5\times10^{-15}$\,s$^{-1}$, about a factor
100 less primarily due to the distance related geometrical
dilution of the incident X-ray flux. The inferred ionisation rates
are much larger than the typical ionisation rates in
Galactic clouds.
Adopting a characteristic density
$n_{H_2}\sim10^{5.5}$\,cm$^{-3}$ from our LVG analysis in Sec.~\ref{sec:others}.
yields
$n/\zeta_{-17} \sim 10$ in the nuclear region of NGC\,3227.
We note that even if the gas density were an order of magnitude higher (leading to proportionally higher $n/\zeta_{-17}$),
the physical conditions will be well within the high-ionisation phase for the molecular chemistry.
\subsubsection{Models of abundance ratio as a function of $n/\zeta_{-17}$}
\begin{figure}
\centering
\includegraphics[angle=0,width=8.8cm]{hcn_co.eps}
\caption{\label{fig:ratio}
Abundance ratio [HCN]/[CO] as a function of $n/\zeta_{-17}$, the ratio of the local gas volume density to ionisation rate,
for 100\,K gas, showing the transition from the high- to low-ionisation phase.
}
\end{figure}
The model computations of \cite{bog05,bog06} show that gas can exist in a high (low) ionisation
phase for small (large) values of $n/\zeta_{-17}$, with a density ratio of $n/\zeta_{-17} \sim 10^3$ marking the cross-over between the two regimes.
The HCN/CO abundance ratio can become large $\gtrsim 10^{-3}$ in the high-ionisation phase even at low gas temperatures, due to the large densities of atomic and ionic carbon.
Here we have re-run these calculations using the same elemental gas-phase abundances, and with a slightly updated reaction set.
Fig.~\ref{fig:ratio} shows the resulting abundance ratios as a function of $n/\zeta_{-17}$
for 100\,K gas.
At very low $n/\zeta_{-17} \lesssim 100$ corresponding to that in the nuclear region of NGC\,3227
the HCN/CO abundance ratio approaches $\sim 10^{-3}$. For comparison, \cite{lep96} find a peak value of
$5\times 10^{-4}$ in their calculation.
These models track the abundances of numerous molecules across many orders of magnitude and show that at low $n/\zeta_{-17}$, the HCN/CO abundance ratio is raised several orders of magnitude above that typically expected, to a level at which it approaches -- and, given the uncertainties in such models, is commensurate with -- that implied by the LVG models.
\cite{har10} have shown that at elevated temperatures $\gtrsim$~300~K, rapid hydrogenation of CN to HCN can increase the HCN/CO abundances further still,
and they comment that such warm gas may be present in the X-ray heated gas near AGN.
We caution however, that their higher temperature
models for which HCN/CO is largest, correspond to $n/\zeta_{-17} = 10^{4.5}$ which is significantly larger than the value we are
invoking as characteristic for the nucleus of NGC\,3227.
The models show that
a combination of X-ray ionisation and heating do yield high HCN abundances, although
not yet quite as high as HCN/CO$\sim10^{-2}$ as
implied by the LVG analysis in Sec.~\ref{sec:others} for the nucleus.
Further chemical modeling is required, but a high HCN/CO intensity ratio
in the nucleus due to elevated X-ray ionisation rates appears plausible.
In this picture, the lower HCN/CO intensity ratio in the ring may simply reflect the lower ionisation rate there,
and not just a lower gas density in this circumgalactic environment.
\section{Summary and Conclusions}
We present an LVG analysis of high-resolution
observations of CO (2-1) and HCN (1-0) line emissions in the central regions of the Seyfert galaxy NGC 3227.
We find that
\begin{itemize}
\item
The HCN(1-0)/CO(2-1) ratio in NGC\,3227 is an order of magnitude higher in the central 80\,pc than in the circumnuclear ring at a radius of 140\,pc.
NGC\,6951 and NGC\,1068 have similarly high published ratios in their central $\sim100$\,pc.
\item
The nuclear HCN/CO intensity ratio in NGC 3227 may reflect optically thick line emission in dense gas with only a weak
constraint on the HCN/CO abundance ratio. However, our
LVG calculations also indicate that the high nuclear ratios in all three of these galaxies are more consistent with
a single set of physical properties corresponding to warm $\sim300$\,K, dense $10^{5.5}$\,cm$^{-3}$ gas,
in which the emission lines are optically thinner, but
in which the HCN/CO abundance ratio is very large $\sim 10^{-2}$.
For these conditions the velocity gradients are $dV/dr \sim 100$\,km\,s$^{-1}$, but would increase significantly at lower temperatures or densities.
Most likely the clouds are gravitationally unbound.
\item
The X-ray ionisation rate at radii less than $\sim 20$~pc may exceed 10$^{-13}$~s$^{-1}$,
and could plausibly lead to high HCN abundances in
molecular gas in the high-ionisation phase where the ratio of the gas density
to the X-ray ionisation rate is small.
\end{itemize}
\begin{acknowledgements}
The authors thank the IRAM staff, in particular Jeremie Boissier and Sascha Trippe, for their invaluable help in obtaining and reducing the data presented in this paper. We thank J.~Graci\'a-Carpio for many useful and interesting discussions. We thank the DFG
for support via German- Israeli Project Cooperation grant STE1869/1-1.GE625/15-1.
\end{acknowledgements}
|
\section{Introduction}
The nature of dark matter and dark energy is one of the most important issues today in physics. There are strong observational evidences in astrophysics and cosmology for the existence of these two components of the cosmic energy budget, indicating that about $95\%$ of the Universe is composed by dark matter (about $25\%$) and by dark energy (about $70\%$), but no direct detection has been reported until now. The usual candidates \textbf{for} dark matter (neutralinos and axions, for example) and dark energy (cosmological constant, quintessence, etc.) lead to very robust scenarios, but at same time they must face theoretical and observational issues. For recent reviews on the subject, see for example \cite{Padmanabhan:2002ji, Peebles:2002gy, Sahni:2004ai, Bertone:2004pz, Sahni:2006pa, Copeland:2006wr, Frieman:2008sn, Martin:2008qp, Caldwell:2009ix, Li:2011sd}.
The strongest issue is perhaps the one regarding dark energy as the vacuum expectation value of some quantum field, which would be a natural candidate, but whose correct theoretical value could be predicted only in the framework of a complete theory of quantum gravity, which still we do not possess. Nevertheless, it is possible, at least, to guess some of the features of this theory. In particular, the holographic principle \cite{'tHooft:1993gx, Susskind:1994vu, Bousso:2002ju} may shed some light on the dark energy problem. According to this principle, in presence of gravity the number of the degrees of freedom of a local quantum system would be related to the area of its boundary, rather than to the volume of the system (as expected when gravity is absent). Following this \textbf{idea}, in \cite{Cohen:1998zx} the authors suggested an entanglement relation between the infrared and ultraviolet \textbf{cutoffs} due to the limitation set by the formation of a black hole, which sets an upper bound for the vacuum energy. We can then interpret the ultraviolet cutoff as the vacuum density value, but still we need an ansatz for the infrared cutoff. As a candidate for such distance, in \cite{Li:2004rb, Huang:2004ai} the authors propose and investigate the future event horizon, tested against type Ia supernovae data and cosmic microwave background anisotropies in \cite{Zhang:2005hs, Zhang:2007sh}. We shall present more detail on this in Sec.~\ref{Sec:HolDE}.
Adding new components of dark energy to the whole energy budget in order to explain the current observation is a way, but not the only one. Since General Relativity has been thoroughly tested up to solar system scales, it may be possible that the Einstein-Hilbert action contain corrections on larger, cosmological, scales thereby candidating as possible explanation of the evolution of the universe. Such modifications should be, in principle, composed by higher order curvature invariant terms (such as $R^2$, $R_{\mu\nu}R^{\mu\nu}$, etc) but also by non-trivial coupling between matter or fields and geometry. See for example \cite{Nojiri:2006ri, Nojiri:2010wj, Amendola:2006kh, Amendola:2006we, Capozziello:2007ec, Sotiriou:2008rp, DeFelice:2010aj} for some reviews on the subject (especially on $f(R)$ theory). It is also worth pointing out that these terms should naturally emerge as quantum corrections in the low energy effective action of quantum gravity or string theory \cite{Buchbinder:1992rb}.
In this paper we connect these two approaches, considering a $f(R,T)$ theory of gravity, where $R$ is the Ricci scalar, whereas $T$ is the trace of the stress-energy momentum. This modified gravity theory has been recently introduced in \cite{Harko:2011kv}, where the authors derived the field equations and considered several cases, relevant in cosmology and astrophysics. As for the former, $f(R,T)$ models have been constructed describing the transition from the matter dominated phase to the late times accelerated one \cite{Houndjo:2011tu}.
Our task here, is to find out which form the function $f(R,T)$ has to have in order to reproduce the same properties of the holographic dark energy proposed in \cite{Li:2004rb}. To this purpose, we employ the same reconstruction scheme proposed and employed in \cite{Capozziello:2005ku, Nojiri:2006gh, Nojiri:2006jy, Nojiri:2006be, Wu:2007tn}. For reference, in order to track the contribution of the $T$ part of the action in the reconstruction, we consider two special $f(R,T)$ models: in the first instance, we investigate the modification $R + 2f(T)$, i.e. the usual Einstein-Hilbert term plus a $f(T)$ correction. In the second instance we consider a $f(R)+\lambda T$ theory, i.e. a $T$ correction to the renown $f(R)$ gravity. In both cases, we consider dark energy accompanied by a pressureless matter component (which would determine $T$).
The paper is organised as follows. In Sec.~\ref{Sec:HolDE}, the equations of motion are established and the holographic dark energy introduced. In Sec.~\ref{Sec:Simpl} and \ref{Sec:ComplCase} the above mentioned cases are analysed. Finally, Sec.~\ref{Sec:DiscConcl} is devoted to discussion and conclusions.
We use $8\pi G = c = 1$ units and adopt the metric formalism, i.e. \textbf{the variation of the action is considered with respect to the metric quantities.}
\section{$f(R,T)$ gravity and holographic Dark Energy}\label{Sec:HolDE}
In \cite{Harko:2011kv}, the following modification of \textbf{Einstein's} theory is proposed:
\begin{equation}\label{actionfRT}
S = \frac{1}{2}\int f(R,T) \sqrt{-g}\;d^4x + \int L_{\rm m} \sqrt{-g}\;d^4x\;,
\end{equation}
where $f(R,T)$ is an arbitrary function of the Ricci scalar $R$ and of the trace $T$ of the energy-momentum tensor, defined as
\begin{equation}
T_{\mu\nu} = -\frac{2}{\sqrt{-g}}\frac{\delta\left(\sqrt{-g}L_{\rm m}\right)}{\delta g^{\mu\nu}}\;,
\end{equation}
where $L_{\rm m}$ is the matter Lagrangian density. We assume the matter lagrangian to depend on the metric, so that
\begin{equation}
T_{\mu\nu} = g_{\mu\nu}L_{\rm m} - 2\frac{\partial L_{\rm m}}{\partial g^{\mu\nu}}\;.
\end{equation}
Varying action \eqref{actionfRT} with respect to the metric $g^{\mu\nu}$, one obtains \cite{Harko:2011kv}
\begin{equation}\label{Eqmod}
f_R(R,T) R_{\mu\nu} - \frac{1}{2}f(R,T)g_{\mu\nu} + \left(g_{\mu\nu}\square - \nabla_\mu\nabla_\nu\right)f_R(R,T) = T_{\mu\nu} - f_T(R,T)T_{\mu\nu} - f_T(R,T)\Theta_{\mu\nu}\;,
\end{equation}
where the \textbf{subscripts} $R$ or $T$ \textbf{imply} derivation with respect that quantity and we have also defined
\begin{equation}
\Theta_{\mu\nu} \equiv g^{\alpha\beta}\frac{\delta T_{\alpha\beta}}{\delta g^{\mu\nu}}\;.
\end{equation}
\textbf{Planning} a cosmological application, we assume matter to be described by a perfect fluid energy-momentum tensor
\begin{equation}
T_{\mu\nu} = \left(\rho + p\right)u_\mu u_\nu - p g_{\mu\nu}\;,
\end{equation}
and that $L_{\rm m} = -p$, so that we have
\begin{equation}
\Theta_{\mu\nu} = -2T_{\mu\nu} - p g_{\mu\nu}\;,
\end{equation}
and Eq.~\eqref{Eqmod} simplifies as
\begin{equation}\label{Eqmod2}
f_R(R,T) R_{\mu\nu} - \frac{1}{2}f(R,T)g_{\mu\nu} + \left(g_{\mu\nu}\square - \nabla_\mu\nabla_\nu\right)f_R(R,T) = T_{\mu\nu} + f_T(R,T)T_{\mu\nu} + p f_T(R,T) g_{\mu\nu}\;.
\end{equation}
In order to compare \textbf{it} with \textbf{Einstein's}, we cast the above equation as follows:
\begin{eqnarray}\label{Eqmod3}
G_{\mu\nu} &=& \frac{1 + f_T(R,T)}{f_R(R,T)}T_{\mu\nu} + \frac{1}{f_R(R,T)}p f_T(R,T) g_{\mu\nu} - \frac{1}{f_R(R,T)}\left(g_{\mu\nu}\square - \nabla_\mu\nabla_\nu\right)f_R(R,T)\nonumber\\ &+& \frac{1}{2f_R(R,T)}f(R,T)g_{\mu\nu} - \frac{1}{2}g_{\mu\nu} R\;,
\end{eqnarray}
where $G_{\mu\nu} \equiv R_{\mu\nu} - Rg_{\mu\nu}/2$ is the Einstein tensor. Now we can identify
\begin{equation}\label{Effmatt}
\tilde{T}_{\mu\nu}^{(m)} = \frac{1 + f_T(R,T)}{f_R(R,T)}T_{\mu\nu} + \frac{1}{f_R(R,T)}p f_T(R,T) g_{\mu\nu}\;,
\end{equation}
as the \textit{effective} matter energy-momentum tensor and
\begin{equation}\label{Effgeom}
\tilde{T}_{\mu\nu}^{(geom)} = - \frac{1}{f_R(R,T)}\left(g_{\mu\nu}\square - \nabla_\mu\nabla_\nu\right)f_R(R,T) + \frac{1}{2f_R(R,T)}f(R,T)g_{\mu\nu} - \frac{1}{2}g_{\mu\nu} R\;,
\end{equation}
as the energy-momentum tensor of a ``geometric'' matter component.
We now assume a \textbf{background} described by the Friedmann-Lema\^{\i}tre-Robertson-Walker metric
\begin{equation}\label{RWmet}
ds^2 = dt^2 - a(t)^2\delta_{ij}dx^idx^j\;,
\end{equation}
with spatially flat hypersurfaces, and find a form for the function $f(R,T)$ which is able to reconstruct \textbf{holographic dark energy}.
\subsection{Holographic Dark Energy}
According to the holographic principle \cite{'tHooft:1993gx, Susskind:1994vu, Bousso:2002ju} an entanglement relation between the infrared (IR) and ultraviolet (UV) cut-offs of a quantum theory, due to the limitation set by the formation of a black hole, sets an upper bound for the vacuum energy \cite{Cohen:1998zx}:
\begin{equation}\label{vacen}
\rho_{\rm v} = \frac{3b^2}{L^2}\;,
\end{equation}
where $b$ is a free parameter and the IR (large scales) cutoff $L$ needs to be specified by an ansatz. We are interested in the one proposed in \cite{Li:2004rb}:
\begin{equation}\label{anshol}
L = R_{\rm h} = a\int_t^\infty \frac{dt'}{a(t')} = a\int_a^\infty \frac{d\bar{a}}{H(\bar{a})\bar{a}^{2}}\;,
\end{equation}
i.e. the future event horizon, that is the distance covered by a photon from now until the remote future. Note that the very presence of a vacuum energy component makes the above integration finite. We consider a model composed by holographic dark energy plus ordinary pressureless matter, i.e.
\begin{equation}\label{Friede}
3H^2 = \rho_{\rm v} + \rho_{\rm m} = \rho_{\rm v} + \rho_{\rm m0}(1 + z)^3\;,
\end{equation}
where $H \equiv \dot{a}/a$ is the Hubble parameter and the dot denotes derivation with respect to the cosmic time. Introduce the critical energy density $\rho_{\rm cr} := 3H^2$\textbf{, we define}
\begin{equation}\label{oliver15}
\Omega_{\rm v} := \frac{\rho_{\rm v}}{\rho_{\rm cr}}=\frac{b^2}{R^2_{\rm h}H^2}\;,
\end{equation}
Using Eqs.\eqref{vacen} and \eqref{anshol}, it is easy to show that
\begin{equation}\label{oliver16}
\dot{R}_{\rm h} = \frac{b}{\sqrt{\Omega_{\rm v}}} - 1\;.
\end{equation}
The holographic dark energy density $\rho_{\rm v}$ evolves according to the conservation law
\begin{equation}\label{oliver17}
\dot{\rho}_{\rm v} + 3H\rho_{\rm v}\left(1 + w_{\rm v}\right) = 0\;,
\end{equation}
because in Eq.~\eqref{Friede} we have implicitly assumed the matter component to conserve separately. Now, using Eqs.~\eqref{vacen}, \eqref{anshol} and \eqref{oliver16}, one can find
\begin{equation}\label{oliver18}
\dot{\rho}_{\rm v} = -\frac{2}{R_{\rm h}}\left(\frac{b}{\sqrt{\Omega_{\rm v}}}-1\right)\rho_{\rm v}\,.
\end{equation}
Comparing \eqref{oliver18} with \eqref{oliver17} one can read off
\begin{equation}\label{oliver19}
w_{\rm v} = -\left(\frac{1}{3} + \frac{2\sqrt{\Omega_{\rm v}}}{3b}\right)\,\,.
\end{equation}
Moreover, combining Eq.~\eqref{anshol} with Eqs.~\eqref{vacen} and \eqref{Friede}, the evolution for $\Omega_{\rm v}$ is determined by the following equation:
\begin{equation}\label{OmegavEvo}
\Omega_{\rm v}' = -\left(1 + \frac{2\sqrt{\Omega_{\rm v}}}{b}\right)\frac{1}{1 + z}\Omega_{\rm v}\left(1 - \Omega_{\rm v}\right)\;,
\end{equation}
where the prime denotes derivation with respect to the redshift $z$. Testing this model against type Ia supernovae and cosmic microwave background anisotropies, $b$ turns out to be constrained around unity, with the case $b < 1$ favoured \cite{Zhang:2005hs, Zhang:2007sh}. Note that, from Eq.~\eqref{oliver19}, $b < 1$ means that the universe will end up in a phantom phase. For more comprehensive analysis of holographic dark energy models, we refer the reader to \cite{Pavon:2005yx, delCampo:2011jp}.
In the next section we investigate a simple case of reconstruction of $f(R,T)$.
\section{A simple case}\label{Sec:Simpl}
We now consider a single perfect fluid model with density $\rho$ and pressure $p$, together with the following ansatz (one of the first considered in \cite{Harko:2011kv}):
\begin{equation}
f(R,T) = R + 2f(T)\;,
\end{equation}
i.e. the action is given by the same Einstein-Hilbert one plus a function of $T$. This is a particularly interesting choice since, from Eqs.~\eqref{Effmatt} and \eqref{Effgeom}, we get
\begin{equation}\label{Tmtil}
\tilde{T}_{\mu\nu} = \left(1 + 2f_T\right)T_{\mu\nu} + 2p f_T g_{\mu\nu} + f(T)g_{\mu\nu}\;.
\end{equation}
For $p = 0$ one has $T = \rho$ and, choosing $f(T) = \lambda T$ one can construct a model with an effective cosmological constant \cite{Poplawski:2006ey}. From Eq.~\eqref{Tmtil} one can read off the effective energy density and pressure of the universe content:
\begin{eqnarray}
\label{rhotot} 3H^2 &=& \rho_{\rm eff} = \left(1 + 2f_T\right)\rho + 2p f_T + f(T)\;,\\
\label{ptot} -2\dot{H} - 3H^2 &=& p_{\rm eff} = p - f(T)\;,
\end{eqnarray}
and therefore a dark energy component may appear, even if we are considering a single perfect fluid model. From Eqs.~\eqref{rhotot} and \eqref{ptot}, it is clear that we can pick out a ``fictitious'' component, due to $f(T)$, described by
\begin{eqnarray}
\rho_{f} &=& 2f_T\rho + 2p f_T + f(T)\;,\\
p_{f} &=& - f(T)\;,
\end{eqnarray}
and, provided $f$ positive, it may well describe a dark energy component, since its pressure is negative. In order to reconstruct the function $f$ starting from the holographic principle, we note that the equation of state parameter of the dark component induced by $f$ is
\begin{equation}
w_{f} = -\frac{f(T)}{2(\rho + p)f_T + f(T)}\;,
\end{equation}
and we identify it with $w_{\rm v}$, the one provided by the holographic dark energy in Eq.~\eqref{oliver19}:
\begin{equation}\label{EqfT}
\frac{f(T)}{2(\rho + p)f_T + f(T)} = \frac{1}{3} + \frac{2\sqrt{\Omega_{\rm v}}}{3b}\;.
\end{equation}
For the standard model given in Eq.~\eqref{Friede}, consider the fluid component to be pressureless matter, i.e. $p = 0$. We are left to solve the following system of equations:
\begin{eqnarray}
\label{fevo} \rho\frac{df(\rho)}{d\rho} &=& f(\rho)\frac{b - \sqrt{\Omega_{\rm v}}}{b + 2\sqrt{\Omega_{\rm v}}}\;,\\
\label{Omevo} \frac{d\Omega_{\rm v}}{d\rho} &=& -\frac{1}{3\rho}\left(1 + \frac{2\sqrt{\Omega_{\rm v}}}{b}\right)\Omega_{\rm v}\left(1 - \Omega_{\rm v}\right)\;,
\end{eqnarray}
where we have used $T = \rho$, because we are considering pressureless matter. In order to solve the above system, we have to fix some initial conditions. Clearly, $\Omega_{\rm v}(\rho = \rho_0) = 1 - \Omega_{\rm m0}$, and we choose $\Omega_{\rm m0} = 0.3$, accordingly with current cosmological observation. For the initial condition on $f$, from Eq.~\eqref{rhotot} (with $p = 0$) we write
\begin{equation}\label{friedeqf0}
\left[1 + 2f_{T}(\rho_0)\right]\Omega_{\rm m0} + \frac{f(\rho_0)}{3H_0^2} = 1\;.
\end{equation}
Evaluating Eq.~\eqref{fevo} today and combining it with Eq.~\eqref{friedeqf0} we find the following algebraic equation determining the initial condition on $f$:
\begin{equation}\label{f0cond}
f(\rho_0)\left(2\frac{b - \sqrt{\Omega_{\rm v0}}}{b + 2\sqrt{\Omega_{\rm v0}}} + 1\right) = 3H_0^2\Omega_{\rm v0}\;.
\end{equation}
As we expected, when $\Omega_{\rm v0} = 0$, then $f(\rho_0) = 0$ and Eq.~\eqref{fevo} implies that $f$ is identically vanishing.\\
\begin{figure}[htbp]
\includegraphics[width=0.45\columnwidth]{Fig1.eps}\;\includegraphics[width=0.45\columnwidth]{Fig2.eps}
\caption{Left panel: evolution of $f(\rho)$. Right panel: evolution of $\Omega_{\rm v}$. The cases considered are $b = 0.6, 0.8, 1.0, 1.2$ (solid black, dashed red, dot-dashed blue and dotted green, respectively). We have chosen as initial conditions in $\Omega_{\rm m0} = 0.3$ the values $\Omega_{\rm v0} = 1 - \Omega_{\rm m0} = 0.7$ and $f_0$ given by Eq.~\eqref{f0cond}. Note that $f$ and $\rho$ are normalised to $3H_0^2$. The vertical lines in the plots represent $\rho = \rho_0$, i.e. the present instant.}
\label{Fig1}
\end{figure}\\
In \figurename{ \ref{Fig1}} we plot the solution of the system of differential equations \eqref{fevo} and \eqref{Omevo}. Note that we normalise $f$ and $\rho$ to $3H_0^2$. As expected from Eq.~\eqref{fevo}, for large values of $\rho$, i.e. far in the past, $f \propto \rho$ because the dark energy component is subdominant. The actual difference among the various choices of $b$ takes place at late times, for small values of $\rho$. Again from inspection of Eq.~\eqref{fevo}, we can see that for large values of $b$ the linear evolution $f \propto \rho$ is again solution. That is why in the left panel of \figurename{ \ref{Fig1}} the curve seems to ``straighten up'' for increasing $b$.
In \figurename{ \ref{Fig2}} we plot the same quantities, but as functions of the redshift, in order to make clearer their cosmological evolution.
\begin{figure}[htbp]
\includegraphics[width=0.45\columnwidth]{Fig3.eps}\;\includegraphics[width=0.45\columnwidth]{Fig4.eps}
\caption{Same as \figurename{ \ref{Fig1}}, but with the redshift $z$ as independent variable.}
\label{Fig2}
\end{figure}
A final remark about the future evolution. It is clear from Eq.~\eqref{fevo}, that when $\Omega_{\rm v} \to 1$, in the remote future, the solution for $f$ gets the asymptotic form
\begin{equation}
f(\rho) \propto \rho^{\frac{b - 1}{b + 2}}\;.
\end{equation}
Therefore, we would have a future singularity for $-2 < b < 1$, as it appears for the relevant cases of \figurename{ \ref{Fig1}} and \figurename{ \ref{Fig2}}. The special case $b = 1$ implies an asymptotically constant $f$.
We now turn our discussion on a more general case, where the curvature $R$ \textbf{comes into} the action \textbf{as} a function to be determined.
\section{A more complicated case}\label{Sec:ComplCase}
Now we turn our attention to the special case
\begin{equation}\label{complcase}
f(R,T) = f(R) + \lambda T\;,
\end{equation}
i.e. a $T$-linear correction to the class of $f(R)$ theories. With the ansatz \eqref{complcase} the matter content \eqref{Effmatt} is ``corrected'' as follows:
\begin{equation}\label{Effmattcomplcase}
\tilde{T}_{\mu\nu}^{(m)} = \frac{1 + \lambda}{f_R(R)}T_{\mu\nu} + \frac{1}{f_R(R)}\lambda p g_{\mu\nu}\;,
\end{equation}
whereas the geometry induced stress-energy tensor is
\begin{equation}\label{Effgeomcomplcase}
\tilde{T}_{\mu\nu}^{(geom)} = - \frac{1}{f_R(R)}\left(g_{\mu\nu}\square - \nabla_\mu\nabla_\nu\right)f_R(R) + \frac{1}{2f_R(R)}[f(R) + \lambda T]g_{\mu\nu} - \frac{1}{2}g_{\mu\nu} R\;.
\end{equation}
Our aim is now to reconstruct the form of the $f(R)$ which is able to reproduce the holographic dark energy paradigm. We again consider a pressureless perfect fluid with density $\rho$ and again assume metric \eqref{RWmet}. From Eqs.~\eqref{Effmattcomplcase} and \eqref{Effgeomcomplcase} the effective density and pressure are the following:
\begin{eqnarray}
\label{rhovcomplcase} 3H^2 &=& \rho_{\rm eff} = \frac{\rho}{f_R} + \frac{3\lambda}{2f_R}\rho - \frac{R}{2} + \frac{f}{2f_R} - 3H\frac{\dot{f}_R}{f_R}\;,\\
\label{pvcomplcase} -2\dot{H} - 3H^2 &=& p_{\rm eff} = \frac{1}{f_R}\left(\ddot{f}_R + 2H\dot{f}_R\right) - \frac{f}{2f_R} - \frac{\lambda}{2f_R}\rho + \frac{R}{2}\;.
\end{eqnarray}
From Eq.~\eqref{rhovcomplcase} it appears that the energy density of the perfect fluid is rescaled by a factor $1/f_R$. Looking at Eq.~\eqref{Friede}, we can extract a form for $\rho_{\rm v}$ in the following way:
\begin{equation}\label{rhovcomplcase2}
\rho_{\rm v} = \frac{\rho}{f_R} - \rho + \frac{3\lambda}{2f_R}\rho - \frac{R}{2} + \frac{f}{2f_R} - 3H\frac{\dot{f}_R}{f_R}\;,
\end{equation}
whereas the form of $p_{\rm v}$ is already given in Eq.~\eqref{pvcomplcase}, since our fluid is pressureless. From Eqs.~\eqref{pvcomplcase} and \eqref{rhovcomplcase2} we can write the following differential equation for $f_R$:
\begin{equation}\label{freqcompl}
\ddot{f}_R - H \dot{f}_R - \left[\rho + \rho_{\rm v}\left(1 + w_{\rm v}\right)\right]f_R = -\rho(1 + \lambda)\;.
\end{equation}
Note, as a cross-check, that for $\rho_{\rm v} = \lambda = 0$ the above equation simplifies to
\begin{equation}\label{freqcompl2simpl}
\ddot{f}_R - H\dot{f}_R - \rho f_R = - \rho\;,
\end{equation}
which possesses the particular solution $f_R = 1$, i.e. $f = R + \Lambda$, the original Einstein-Hilbert action plus an integration ``cosmological'' constant. We expect this solution to be the only one, otherwise there would exist an alternative $f(R)$ theory which would behave exactly as general relativity. Let us speculate a bit more on this point. If $\rho_{\rm v} = \lambda = 0$, i.e. for a pure Einstein-de Sitter universe, we have from Eq.~\eqref{Friede}
\begin{equation}
H = \frac{2}{3t}\;, \qquad \rho = \rho_0\frac{t^2_0}{t^2}\;,
\end{equation}
where $t_0$ is the present cosmic time (i.e. the age of the universe). Considering the homogeneous part of Eq.~\eqref{freqcompl2simpl} and looking for a solution of the form $f_R \propto t^n$, we find:
\begin{equation}\label{freqcompl2simplhom}
n(n - 1) - \frac{2}{3}n - \rho_0 t_0^2 = 0\;,
\end{equation}
which gives:
\begin{equation}\label{freqcompl2simplhomsol}
n_{1,2} = \frac{5}{6} \pm \sqrt{\frac{25}{36} + \rho_0 t_0^2}\;,
\end{equation}
and the general solution of Eq.~\eqref{freqcompl2simpl} can be written as:
\begin{equation}\label{freqcompl2simplhomsol2}
f_R = 1 + C_1\;t^{n_1} + C_2\;t^{n_2}\;.
\end{equation}
Now, the initial conditions we adopt here are $f_R(t_0) = 1$ and $\dot{f}_R(t_0) = 0$. The reason is essentially not spoiling the agreement between general relativity and solar system tests, see \cite{Capozziello:2005ku, Nojiri:2006gh, Nojiri:2006jy, Nojiri:2006be, Wu:2007tn}. However, we stress here that these two conditions also imply $C_1 = C_2 = 0$ and therefore restore the general relativity limit $f_R = 1$ when $\rho_{\rm v} = \lambda = 0$.
Changing the variable to the redshift and employing Eqs.~\eqref{Friede} and \eqref{OmegavEvo} one can recast Eq.~\eqref{freqcompl} in the following compact form:
\begin{eqnarray}\label{freqcompl2}
(1 + z)^2f''_R + \frac{1 + z}{2}\left[7 - \Omega_{\rm v}\left(1 + \frac{2\sqrt{\Omega_{\rm v}}}{b}\right)\right]f_R' - 3\left[1 - \Omega_{\rm v}\left(\frac{1}{3} + \frac{2\sqrt{\Omega_{\rm v}}}{3b}\right)\right]f_R = \nonumber\\ - 3(1 + \lambda)(1 - \Omega_{\rm v})\;,
\end{eqnarray}
where again the prime denotes derivation with respect to the redshift $z$. The curvature $R$ can be easily found as
\begin{equation}
R = -6(\dot{H} + 2H^2) = -\frac{3H_0^2\Omega_{\rm m0}(1 + z)^3}{1 - \Omega_{\rm v}}\left[1 + \Omega_{\rm v}\left(1 + \frac{2\sqrt{\Omega_{\rm v}}}{b}\right)\right]\;.
\end{equation}
The initial conditions $f_R(t_0) = 1$ and $\dot{f}_R(t_0) = 0$, translated to the redshift variable, are
\begin{eqnarray}
\left.\frac{d^2f}{dz^2}\right|_{z = 0} = \left.\frac{d^2R}{dz^2}\right|_{z = 0}\;, \qquad \left.\frac{df}{dz}\right|_{z = 0} = \left.\frac{dR}{dz}\right|_{z = 0}\;,\\
\end{eqnarray}
Finally, the initial condition on $f$ can be extracted by Eq.~\eqref{rhovcomplcase}, being that $\rho_{\rm v0} = 3H_0^2 - \rho_{\rm m0}$. Thus, we have
\begin{equation}
f(z = 0) = R(z = 0) + 6H_0^2\left(1 - \Omega_{\rm m0} - \frac{3}{2}\lambda\Omega_{\rm m0}\right)\;.
\end{equation}
Note the correction due to the $\lambda T$ term.
\begin{figure}[htbp]
\includegraphics[width=0.45\columnwidth]{Fig5.eps}\;\includegraphics[width=0.45\columnwidth]{Fig6.eps}
\caption{Evolution of $f$ as a function of the curvature $R$. Left panel: $\lambda = 0$, i.e. we are considering a pure $f(R)$ theory, and $b = 0.6, 0.8, 1.0, 1.2$ (solid black, dashed red, dot-dashed blue and dotted green, respectively). Right panel: $b = 1.0$ and $\lambda = -0.2, 0, 0.2, 0.4$ (solid black, dashed red, dot-dashed blue and dotted green, respectively. The curves appear superposed.). We have chosen as initial conditions in $\Omega_{\rm m0} = 0.3$ the values $\Omega_{\rm v0} = 1 - \Omega_{\rm m0} = 0.7$. Note that $f$ and $R$ are normalised to $3H_0^2$ and the redshift interval chosen is $0 < z < 10$.}
\label{Fig3}
\end{figure}
In \figurename{ \ref{Fig3}} we plot the solution for $f$. Note that we normalise $f$ and $R$ to $3H_0^2$. In the left panel we fix $\lambda = 0$, i.e. we are actually considering a pure $f(R)$ theory, and vary $b$. In the right panel, on the other hand, we consider positive and negative values of $\lambda$. As one may note, $\lambda$ has a poor influence on the evolution of $f$. We could expect this from inspection of Eq.~\eqref{freqcompl2}. Indeed, $\lambda$ only enters the source term on the right hand side and therefore, when $\Omega_{\rm v}$ grows to unity, its impact on the evolution of $f$ is weak. On the other hand, larger values of $\lambda$ may have a relevant effect at early times, determining the slope of $f$.
\begin{figure}[ht]
\includegraphics[width=0.45\columnwidth]{Fig7.eps}\;\includegraphics[width=0.45\columnwidth]{Fig8.eps}
\caption{Same as \figurename{ \ref{Fig3}}, but with $-0.9 < z < 1$. Note, in the left panel, that the evolution of $f$ starts from below for all the cases.}
\label{Fig4}
\end{figure}
\newpage
In \figurename{ \ref{Fig4}} and \figurename{ \ref{Fig5}} we display the future evolution of $f$, both as function of $R$ or of $z$. For the same reason stated above, the effect of $\lambda$ is not relevant.
\begin{figure}[ht]
\includegraphics[width=0.45\columnwidth]{Fig9.eps}\;\includegraphics[width=0.45\columnwidth]{Fig10.eps}
\caption{Same as \figurename{ \ref{Fig4}}, but with $f$ as function of $z$.}
\label{Fig5}
\end{figure}
\newpage
Finally, in \figurename{ \ref{Fig6}} we plot the solution for $\Omega_{\rm v}$. Again, its evolution appears to be independent of $\lambda$.
\begin{figure}[ht]
\includegraphics[width=0.45\columnwidth]{Fig11.eps}\;\includegraphics[width=0.45\columnwidth]{Fig12.eps}
\caption{Evolution of $\Omega_{\rm v}$ as a function of the redshift. Left panel: $\lambda = 0$, i.e. we are considering a pure $f(R)$ theory, and $b = 0.6, 0.8, 1.0, 1.2$ (black, dashed red, dash-dotted blue and dotted green, respectively). Right panel: $b = 1.0$ and $\lambda = -0.2, 0, 0.2, 0.4$ (black, dashed red, dash-dotted blue and dotted green, respectively). Note that the curves are superposed. We have chosen as initial conditions in $\Omega_{\rm m0} = 0.3$ the values $\Omega_{\rm v0} = 1 - \Omega_{\rm m0} = 0.7$.}
\label{Fig6}
\end{figure}
\newpage
\section{Discussion and Conclusions}\label{Sec:DiscConcl}
In this work we have investigated a description of holographic dark energy in terms of suitably reconstructed $f(R, T)$ gravity theories. The latter have been recently introduced as modifications of Einstein's theory possessing some interesting solutions \textbf{which are} relevant in cosmology and astrophysics \cite{Harko:2011kv}.
We have \textbf{considered two special} types of models: $f(R,T) = R + 2f(T)$, i.e. a correction to the Einstein-Hilbert action depending on the matter content, and $f(R,T) = f(R) + \lambda T$, i.e. a simple $T$-linear correction to the class of $f(R)$ theories.
Since we have assumed the matter content to be a pressureless perfect fluid, then $T = \rho$, i.e. the corrections assumed are directly dependent on the energy density of the universe content. We \textbf{have} constructed differential equations for the function $f$ under investigation and numerically solved \textbf{them}, physically specifying the required initial conditions. Our simple analysis shows that holographic dark energy models are contained in the larger class of $f(R,T)$ theories, at least considering a given background evolution of the universe.
It would be interesting to investigate how the evolution of matter perturbations would change, depending on the description of dark energy. We expect, in principle, different results when using holographic dark energy or its $f(R,T)$ reconstruction and therefore there is possibility for discriminating between the two descriptions. For example, it would be interesting to adapt the recently proposed scheme for perturbations in $f(R)$ theories \cite{Bertacca:2011wu} to the broader $f(R,T)$ class. We leave this as a future work.
\section*{Acknowledgements}
MJSH and OFP thank Professor S. D. Odintsov for useful comments and also CNPq (Brazil) for partial financial support.
\bibliographystyle{unsrt}
|
\section{Introduction}
All graphs in this paper are undirected, finite and simple. We refer
to book \cite{bondy} for notation and terminology not described
here. A path $u=u_1,u_2,\ldots,u_k=v$ is called a $P_{u,v}$ path.
Denote by $u_iPu_j$ the subpath $u_i,u_{i+1},\ldots,u_j$ for $i\leq
j$. The $length\ \ell(P)$ of a path $P$ is the number of edges in
$P$. The $distance$ between two vertices $x$ and $y$ in $G$, denoted
by $d_{G}(x,y)$, is the length of a shortest path between them. The
$eccentricity$ of a vertex $x$ in $G$ is $ecc_{G}(x)=max_{y\in
V(G)}d(x,y)$. The $radius$ and $diameter$ of $G$ are
$rad(G)=min_{x\in V(G)}ecc(x)$ and $diam(G)=max_{x\in V(G)}ecc(x)$,
respectively. A vertex $u$ is a $center$ of a graph $G$ if
$ecc(u)=rad(G)$. The oriented diameter of a bridgeless graph $G$ is
$\min\{\,diam(H)\ |$ $H\ is\ an\ orientation\ of\ G\}$, and the
oriented radius of a bridgeless graph $G$ is $\min\{\,rad(H)\ |\, H\
is\ an\ orientation\ of\ G\}$. For any graph $G$ with
edge-connectivity $\lambda(G)=0,1$, $G$ has oriented radius (resp.
diameter) $\infty$.
In 1939, Robbins solved the One-Way Street Problem and proved that a
graph $G$ admits a strongly connected orientation if and only if $G$
is bridgeless, that is, $G$ does not have any cut-edge. Naturally,
one hopes that the oriented diameter of a bridgeless graph is as
small as possible. Bondy and Murty suggested to study the
quantitative variations on Robbins' theorem. In particular, they
conjectured that there exists a function $f$ such that every
bridgeless graph with diameter $d$ admits an orientation of diameter
at most $f(d)$.
In 1978, Chv\'atal and Thomassen \cite{chv} obtained some general
bounds.
\begin{theorem}[\bf Chv\'atal and Thomassen 1978 \cite{chv}]\label{T1}
For every bridgeless graph $G$, there exists an orientation $H$ of
$G$ such that
\[rad(H)\leq rad(G)^2+rad(G),\]
\[diam(H)\leq 2rad(G)^2+2rad(G).\] Moreover, the above bounds are
optimal.
\end{theorem}
There exists a minor error when they constructed the graph $G_d$
which arrives at the upper bound when $d$ is odd. Kwok, Liu and West
gave a slight correction in \cite{kwo}.
They also showed that determining whether an arbitrary graph can be
oriented so that its diameter is at most 2 is NP-complete. Bounds
for the oriented diameter of graphs have also been studied in terms
of other parameters, for example, radius, dominating number
\cite{chv,fom,kwo,sol}, etc. Some classes of graphs have also been
studied in \cite{fom, koh1,koh2,kon,mcc}.
Let $\eta(G)$ be the smallest integer such that every edge of $G$
belongs to a cycle of length at most $\eta(G)$. In this paper, we
show the following result.
\begin{theorem}\label{T2}
For every bridgeless graph $G$, there exists an orientation $H$ of
$G$ such that
\[rad(H)\leq \sum_{i=1}^{rad(G)}\min\{2i,\eta(G)-1\}\leq
rad(G)(\eta(G)-1),\]
\[diam(H)\leq 2\sum_{i=1}^{rad(G)}\min\{2i,\eta(G)-1\}\leq
2rad(G)(\eta(G)-1).\]
\end{theorem}
Note that $\sum_{i=1}^{rad(G)}\min\{2i,\eta(G)-1\}\leq
rad(G)^2+rad(G)$ and $diam(H)\leq 2rad(G)$. So our result implies
Chv\'atal and Thomassen's Theorem \ref{T1}.
A path in an edge-colored graph $G$, where adjacent edges may have
the same color, is called $rainbow$ if no two edges of the path are
colored the same. An edge-coloring of a graph $G$ is a $rainbow\
edge$-$coloring$ if every two distinct vertices of graph $G$ are
connected by a rainbow path. The $rainbow\ connection\ number\
rc(G)$ of $G$ is the minimum integer $k$ for which there exists a
rainbow $k$-edge-coloring of $G$. It is easy to see that
$diam(G)\leq rc(G)$ for any connected graph $G$. The rainbow
connection number was introduced by Chartrand et al. in \cite{char}.
It is of great use in transferring information of high security in
multicomputer networks. We refer the readers to \cite{chak} for
details.
Chakraborty et al. \cite{chak} investigated the hardness and
algorithms for the rainbow connection number, and showed that given
a graph $G$, deciding if $rc(G)=2$ is $NP$-complete. Bounds for the
rainbow connection number of a graph have also been studies in terms
of other graph parameters, for example, radius, dominating number,
minimum degree, connectivity, etc. \cite{bas,char,kri}. Cayley
graphs and line graphs were studied in \cite{liliu} and \cite{lis},
respectively.
A subgraph $H$ of a graph $G$ is called $isometric$ if the distance
between any two distinct vertices in $H$ is the same as their
distance in $G$. The size of a largest isometric cycle in $G$ is
denoted by $\zeta(G)$. Clearly, every isometric cycle is an induced
cycle and thus $\zeta(G)$ is not larger than the chordality, where
$chordality$ is the length of a largest induced cycle in $G$. In
\cite{bas}, Basavaraju, Chandran, Rajendraprasad and Ramaswamy got
the the following sharp upper bound for the rainbow connection
number of a bridgeless graph $G$ in terms of $rad(G)$ and
$\zeta(G)$.
\begin{theorem}[\bf Basavaraju et al. \upshape\cite{bas}]\label{T3}
For every bridgeless graph $G$,
\[rc(G)\leq\sum_{i=1}^{rad(G)}\min\{2i+1,\zeta(G)\}\leq rad(G)\zeta(G).\]
\end{theorem}
In this paper, we show the following result.
\begin{theorem} \label{T4}
For every bridgeless graph $G$,
\[rc(G)\leq \sum_{i=1}^{rad(G)}\min\{2i+1,\eta(G)\}\leq
rad(G)\eta(G).\]
\end{theorem}
From Lemma~\ref{L2} of Section~2, we will see that $\eta(G)\leq
\zeta(G)$. Thus our result implies Theorem~\ref{T3}.
This paper is organized as follows: in Section $2$, we introduce
some new definitions and show several lemmas. In Section $3$, we
prove Theorem~\ref{T2} and study upper for the oriented radius
(resp. diameter) of plane graphs, edge-transitive graphs and general
(bipartite) graphs. In Section $4$, we prove Theorem~\ref{T4} and
study upper for the rainbow connection number of plane graphs,
edge-transitive graphs and general (bipartite) graphs.
\section{Preliminaries}
In this section, we introduce some definitions and show several
lemmas.
\begin{definition}\upshape For any $x\in V(G)$ and $k\geq 0$,
the $k$-$step\ open\ neighborhood$ is $\{y\,|\,d(x,y)=k\}$ and
denoted by $N_k(x)$, the $k$-$step\ closed\ neighborhood$ is
$\{y\,|\,d(x,y)\linebreak[3]\leq k\}$ and denoted by $N_k[x]$. If
$k=1$, we simply write $N(x)$ and $N[x]$ for $N_1(x)$ and $N_1[x]$,
respectively.\end{definition}
\begin{definition}\upshape
Let $G$ be a graph and $H$ be a subset of $V(G)$ (or a subgraph of
$G$). The edges between $H$ and $G\setminus H$ are called $legs$ of
$H$. An $H$-$ear$ is a path $P=(u_0,u_1,\ldots,u_k)$ in $G$ such
that $V(H)\cap V(P)=\{u_0,u_k\}$. The vertices $u_0,\,u_k$ are
called the $foot$ of $P$ in $H$ and $u_0u_1,\,u_{k-1}u_k$ are called
the $legs$ of $P$. The $length$ of an $H$-ear is the length of the
corresponding path. If $u_0=u_k$, then $P$ is called a $closed\
H$-$ear$. For any leg $e$ of $H$, denote by $\ell(e)$ the smallest
number such that there exists an $H$-ear of length $\ell(e)$
containing $e$, and such an $H$-ear is called an $optimal$
$(H,e)$-ear.
\end{definition}
Note that for any optimal $(H,e)$-ear $P$ and every pair $(x,y)\neq
(u_0,u_k)$ of distinct vertices of $P$, $x$ and $y$ are adjacent on
$P$ if and only if $x$ and $y$ are adjacent in $G$.
\begin{definition}\upshape
For any two paths $P$ and $Q$, the joint of $P$ and $Q$ are the
common vertex and edge of $P$ and $Q$. Paths $P$ and $Q$ have $k\
continuous\ common\ segments$ if the common vertex and edge are $k$
disjoint paths.
\end{definition}
\begin{definition}\upshape
Let $P$ and $Q$ be two paths in $G$. Call $P$ and $Q$ $independent$
if they has no common internal vertex.
\end{definition}
\begin{lemma}\label{L1}
Let $n\geq 1$ be an integer, and let $G$ be a graph, $H$ be a
subgraph of $G$ and $e_i=u_iv_i$ be a leg of $H$ and
$P_i=P_{u_iw_i}$ be an optimal $(G,e_i)$-ear, where $1\leq i\leq n$
and $u_i,w_i$ are the foot of $P_i$. Then for any leg
$e_j=u_jv_j\neq e_i,\,1\leq i\leq n$, either there exists an optimal
$(H,e_j)$-ear $P_j=P_{u_jw_j}$ such that either $P_i$ and $P_j$ are
independent for any $P_i,\,1\leq i\leq n$, or $P_i$ and $P_j$ have
only one continuous common segment containing $w_j$ for some $P_i$.
\end{lemma}
\begin{proof} Let $P_j$ be an optimal $(H,e_j)$-ear. If $P_i$ and $P_j$
are independent for any $i$, then we are done. Suppose that $P_i$
and $P_j$ have $m$ continuous common segments for some $i$, where
$m\geq 1$.
\begin{figure}[h,t,b]
\begin{center}
\scalebox{0.7}[0.7]{\includegraphics{Fig1.eps}}\vspace{0.5cm}
Figure 1. Two $H$-ears $P_i$ and $P_j$.
\end{center}
\end{figure}
When $m\geq 2$, we first construct an optimal $(H,e_j)$-ear $P^*_j$
such that $P_i$ and $P^*_j$ has only one continuous common segment.
Let $P_{i_1},P_{i_2},\ldots,P_{i_m}$ be the $m$ continuous common
segments of $P_i$ and $P_j$ and they appear in $P_i$ in that order.
See Figure~1 for details. Furthermore, suppose that $x_{i_k}$ and
$y_{i_k}$ are the two ends of the path $P_{i_k}$ and they appear in
$P_i$ successively. We say that the following claim holds.
\noindent{\itshape Claim 1:
$\ell(y_kP_ix_{k+1})=\ell(y_kP_jx_{k+1})$ for any $1\leq k\leq
m-1$.}
If not, that is, there exists an integer $k$ such that
$\ell(y_kP_ix_{k+1})\neq\ell(y_kP_jx_{k+1})$. Without loss of
generality, we assume $\ell(y_kP_ix_{k+1})<\ell(y_kP_jx_{k+1})$.
Then we shall get a more shorter path $H$-ear containing $e_j$ by
replacing $y_kP_jx_{k+1}$ with $y_kP_ix_{k+1}$, a contradiction.
Thus $\ell(y_kP_ix_{k+1})=\ell(y_kP_jx_{k+1})$ for any $k$.
Let $P^*_j$ be the path obtained from $P_j$ by replacing
$y_kP_jx_{k+1}$ with $y_kP_ix_{k+1}$, and let $P_j=P^*_j$. If the
continuous common segment of $P_i$ and $P_j$ does not contain $w_j$.
Suppose $x$ and $y$ are the two ends of the common segment such that
$x$ and $y$ appeared on $P$ starting from $u_i$ to $w_i$
successively. Similar to Claim~1, $\ell(yP_iw_i)=\ell(yP_jw_j)$. Let
$P^*_j$ be the path obtained from $P_j$ by replacing $yP_jw_j$ with
$yP_iw_i$. Clearly, $P^*_j$ is our desired optimal $(H,u_jv_j)$-ear.
\end{proof}
\begin{lemma}\label{L2}
For every bridgeless graph $G$, $\eta(G)\leq\zeta(G)$.
\end{lemma}
\begin{proof}
Suppose that there exists an edge $e$ such that the length $\ell(C)$
of the smallest cycle $C$ containing $e$ is larger than $\zeta(G)$.
Then, $C$ is not an isometric cycle since the length of a largest
isometric cycle is $\zeta(G)$. Thus there exist two vertices $u$ and
$v$ on $C$ such that $d_{G}(u,v)<d_{C}(u,v)$. Let $P$ be a shortest
path between $u$ and $v$ in $G$. Then a closed trial $C'$ containing
$e$ is obtained from the segment of $C$ containing $e$ between $u$
and $v$ by adding $P$. Clearly, the length $\ell(C')$ is less than
$\ell(C)$. We can get a cycle $C''$ containing $e$ from $C'$. Thus
there exists a cycle $C''$ containing $e$ with length less than
$\ell(C)$, a contradiction. Therefore $\eta(G)\leq\zeta(G)$.
\end{proof}
\begin{lemma}\label{L3}
Let $G$ be a bridgeless graph and $u$ be a center of $G$. For any
$i\leq rad(G)-1$ and every leg $e$ of $N_i(u)$, there exists an
optimal $(N_i[u],e)$-ear with length at most
$\min\{2(rad(G)-i)+1,\eta(G)\}$.
\end{lemma}
\begin{proof}
Let $P$ be an optimal $(N_i[u],e)$-ear. Since $e$ belongs a cycle
with length at most $\eta(G)$, $\ell(P)\leq \eta(G)$. On the other
hand, if $\ell(P)\geq 2(rad(G)-i)+1$, then the middle vertex of $P$
has length at least $rad(G)-i+1$ from $N_i[u]$, a contradiction.
\end{proof}
\section{Oriented diameter}
At first, we have the following observation.
\begin{observation}
Let $G$ be a graph and $H$ be a bridgeless spanning subgraph of $G$.
Then the oriented radius (resp. diameter) of $G$ is not larger than
the oriented radius (resp. diameter) of $H$.
\end{observation}
\noindent{\itshape Proof of Theorem~\ref{T2}:} We only need to show
that $G$ has an orientation $H$ such that $rad(H)\leq
\sum_{i=1}^{rad(G)}\min\{2i,\eta(G)-1\}\leq rad(G)(\eta(G)-1)$. Let
$u$ be a center of $G$ and let $H_0$ be the trivial graph with
vertex set $\{u\}$. We assert that {\itshape there exists a subgraph
$G_i$ of $G$ such that $N_i[u]\subseteq V(G_i)$ and $G_i$ has an
orientation $H_i$ satisfying that $rad(H_i)\leq ecc_{H_i}(u)\leq
\Sigma_{j=1}^i\min\{2(rad(G)-j),\eta(G)-1\}$.}
{\itshape Basic step:} When $i=1$, we omit it since the proof of
this step is similar to that of the following induction step.
{\itshape Induction step:} Assume that the above assertion holds for
$i-1$. Next we show that the above assertion also holds for $i$. For
any $v\in N_i(u)$, either $v\in V(H_{i-1})$ or $v\in N(H_i)$ since
$N_{i-1}[u]\subseteq V(H_{i-1})$. If $N_i(u)\subseteq V(H_{i-1})$,
then let $H_i=H_{i-1}$ and we are done. Thus, we suppose
$N_i(u)\not\subseteq V(H_{i-1})$ in the following.
Let $X=N_i(u)\setminus V(H_{i-1})$. Pick $x_1\in X$, let $y_1$ be a
neighbor of $x_1$ in $H_{i-1}$ and let $P_1=P_{y_1z_1}$ be an
optimal $(H_{i-1},x_1y_1)$-ear. We orient $P$ such that $P_1$ is a
directed path. Pick $x_2\in X$ satisfying that all incident edges of
$x_2$ are not oriented. Let $y_2$ be a neighbor of $x_2$ in
$H_{i-1}$. If there exists an optimal $(H_{i-1},x_2y_2)$-ear $P_2$
such that $P_1$ and $P_2$ are independent, then we can orient $P_2$
such that $P_2$ is a directed path. Otherwise, by Lemma~\ref{L1}
there exists an optimal $(H_{i-1},x_2y_2)$-ear $P_2=P_{y_2z_2}$ such
that $P_1$ and $P_2$ has only one continuous common segment
containing $z_2$. Clearly, we can orient the edges in
$E(P_2)\setminus E(P_1)$ such that $P_2$ is a directed path. We can
pick the vertices of $X$ and oriented optimal $H$-ears similar to
the above method until that for any $x\in X$, at least two incident
edges of $x$ are oriented. Let $H_i$ be the graph obtained from
$H_{i-1}$ by adding vertices in $V(G)\setminus V(H_{i-1})$, which
has at least two new oriented incident edges, and adding new
oriented edges. Clearly, $N_i[u]\subseteq V(H_i)=V(G_i)$.
Now we show that $rad(H_i)\leq
\Sigma_{j=1}^i\min\{2(rad(G)-i),\eta(G)-1\}$. It suffices to show
that for every vertex $x$ of $H_i$, $d_{H_i}(H_{i-1},x)\leq
\min\{2(rad(G)-i),\eta(G)-1\}$ and $d_{H_i}(x,H_{i-1})\leq
\min\{2(rad(G)-i),\eta(G)-1\}$. If $x\in V(H_{i-1})$, then the
assertion holds by inductive hypothesis. If $x\not\in V(H_{i-1})$.
Let $P$ be a directed optimal $(H_i,e)$-ear containing $x$, where
$e$ is some leg of $H_{i-1}$ (such a leg and such an ear exists by
the definition of $H_i$. By Lemma~\ref{L3}, $\ell(P)\leq
\min\{2(rad(G)-i)+1,\eta(G)\}$. Thus, $d_{H_i}(x,H_{i-1})\leq
\min\{2(rad(G)-i),\eta(G)-1\}$ and $d_{H_i}(H_{i-1},x)\leq
\min\{2(rad(G)-i),\eta(G)-1\}$. Therefore, $rad(H_i)\leq
\Sigma_{j=1}^i\min\{2(rad(G)-j),\eta(G)-1\}$. \hfill$\Box$
\begin{remark}
The above theorem is optimal since it implies Chv\'atal and
Thomassen's optimal Theorem \ref{T1}. Readers can see \cite{chv,
kwo} for optimal examples.
\end{remark}
The following example shows that our result is better than that of
Theorem~\ref{T1}.
\begin{example}\upshape
Let $H_3$ be a triangle with one of its vertices designated as root.
In order to construct $H_r$, take two copies of $H_{r-1}$. Let $H_r$
be the graph obtained from the triangle $u_0,u_1,u_2$ by identifying
the root of first (resp. second) copy of $H_{r-1}$ with $u_1$ (resp.
$u_2$), and $u_0$ be the root of $H_r$. Let $G_r$ be the graph
obtained by taking two copies of $H_r$ and identifying their roots.
See Figure~2 for details. It is easy to check that $G_r$ has radius
$r$ and every edge belongs to a cycle of length $\eta(G)=3$. By
Theorem~\ref{T1}, $G_r$ has an orientation $H_r$ such that
$rad(H_r)\leq r^2+r$ and $diam(H_r)\leq 2r^2+2r$. But, by
Theorem~\ref{T2}, $G_r$ has an orientation $H_r$ such that
$rad(G)\leq 2r$ and $diam(G)\leq 4r$. On the other hand, it is easy
to check that all the strong orientations of $G_r$ has radius $2r$
and diameter $4r$.
\end{example}
\begin{figure}[h,t,b]
\begin{center}
\scalebox{0.8}[0.8]{\includegraphics{Fig2.eps}}
\vspace{0.5cm} Figure~2. The graph $G_3$ which has oriented\\ radius
$6$
and oriented diameter $12$.
\end{center}
\end{figure}
We have the following result for plane graphs.
\begin{theorem}
Let $G$ be a plane graph. If the length of the boundary of every
face is at most $k$, then $G$ has an oriented $H$ such that
$rad(H)\leq rad(G)(k-1)$ and $diam(H)\leq 2rad(G)(k-1)$.
\end{theorem}
Since every edge of a maximal plane (resp. outerplane) graph belongs
to a cycle with length $3$, the following corollary holds.
\begin{corollary}
Let $G$ be a maximal plane (resp. outerplane) graph. Then there
exists an orientation $H$ of $G$ such that $rad(H)\leq 2rad(G)$ and
$rad(H)\leq 4rad(G)$.
\end{corollary}
A graph $G$ is $edge$-$transitive$ if for any $e_1,e_2\in E(G)$,
there exists an automorphism $g$ such that $g(e_1)=e_2$. We have the
following result for edge-transitive graphs.
\begin{theorem}
Let $G$ be a bridgeless edge-transitive graph. Then $G$ has an
orientation $H$ such that $rad(H)\leq rad(G)(g(G)-1)$ and
$diam(H)\leq 2rad(G)(g(G)-1)$, where $g(G)$ is the girth of $G$,
that is, the length of a smallest induced cycle.
\end{theorem}
For general bipartite graphs, the following theorem holds.
\begin{theorem}\label{T5}
Let $G=(V_1\cup V_2,E)$ be a bipartite graph with $|V_1|=n$ and
$|V_2|=m$. If $d(x)\geq k>\lceil m/2 \rceil$ for any $x\in V_1$,
$d(y)\geq r>\lceil n/2 \rceil$ for any $y\in V_2$, then there exists
an orientation $H$ of $G$ such that $rad(H)\leq 9$.
\end{theorem}
\begin{proof}
It suffices to show that $rad(G)\leq 3$ and $\eta(G)\leq 4$ by
Theorem~\ref{T2}.
First, we show that $rad(G)\leq 3$. Fix a vertex $x$ in $G$, and let
$y$ be any vertex different from $x$ in $G$. If $x$ and $y$ belong
to the same part, without loss of generality, say $x,y\in V_1$. Let
$X$ and $Y$ be neighborhoods of $x$ and $y$ in $V_2$, respectively.
If $X\cap Y=\emptyset$, then $|V_2|\geq |X|+|Y|\geq 2k>m$, a
contradiction. Thus $X\cap Y\neq \emptyset$, that is, there exists a
path between $x$ and $y$ of length two. If $x$ and $y$ belong to
different parts, without loss of generality, say $x\in V_1,y\in
V_2$. Suppose $x$ and $y$ are nonadjacent, otherwise there is
nothing to do. Let $X$ and $Y$ be neighborhoods of $x$ and $y$ in
$G$, and let $X'$ be the set of neighbors except for $x$ of $X$ in
$G$. If $X'\cap Y=\emptyset$, then $|V_1|\geq 1+|Y|+|X'|\geq
1+r+(r-1)=2r>n$, a contradiction (Note that $|X'|\geq r-1$). Thus
$X'\cap Y\neq\emptyset$, that is, there exists a path between $x$
and $y$ of length three in $G$.
Next we show that $\eta(G)\leq 4$. Let $xy$ be any edge in $G$. Let
$X$ be the set of neighbors of $x$ except for $y$ in $G$, let $Y$ be
the set of neighbors of $y$ except for $x$ in $G$, let $X'$ be the
set of neighbors except for $x$ of $X$ in $G$. If $X'\cap
Y=\emptyset$, then $|V_1|\geq 1+|Y|+|X'|\geq 1+(r-1)+(r-1)=2r-1>n$,
a contradiction (Note that $|X'|\geq r-1$). Thus $X'\cap
Y\neq\emptyset$, that is, there exists a cycle containing $xy$ of
length four in $G$.
\end{proof}
\begin{remark}
The degree condition is optimal. Let $m, n$ be two even numbers with
$n,m\geq 2$. Since $K_{n/2,m/2}\cup K_{n/2,m/2}$ is disconnected,
the oriented radius (resp. diameter) of $K_{n/2,m/2}\cup
K_{n/2,m/2}$ is $\infty$.
\end{remark}
For equal bipartition $k$-regular graph, the following corollary
holds.
\begin{corollary}
Let $G=(V_1\cup V_2, E)$ be a $k$-regular bipartite graph with
$|V_1|=|V_2|=n$. If $k> n/2$, then there exists an orientation $H$
of $G$ such that $rad(H)\leq 9$.
\end{corollary}
The following theorem holds for general graphs.
\begin{theorem}
Let $G$ be a graph.
$(i)$ If there exists an integer $k\geq 2$ such that
$|N_k(u)|>n/2-1$ for every vertex $u$ in $G$, then $G$ has an
orientation $H$ such that $rad(H)\leq 4k^2$ and $diam(H)\leq 8k^2$.
$(ii)$ If $\delta(G)>n/2$, then $G$ has an orientation $H$ such that
$rad(H)\leq 4$ and $diam(H)\leq 8$.
\end{theorem}
\begin{proof}
Since methods of proofs of $(i)$ and $(ii)$ are similar, we only
prove $(i)$. For $(i)$, it suffices to show that $rad(G)\leq 2k$ and
$\eta(G)\leq 2k+1$ by Theorem~\ref{T2}.
We first show $rad(G)\leq 2k$. Fix $u$ in $G$, for every $v\in
V(G)$, if $v\in N_k[u]$, then $d(u,v)\leq k$. Suppose $v\not\in
N_k[u]$, we have $N_k(u)\cap N_k(v)\neq\emptyset$. If not, that is,
$N_k(u)\cap N_k(v)=\emptyset$, then $|N_k(u)|+|N_k(v)|+2>n$ (a
contradiction). Thus $d(u,v)\leq 2k$.
Next we show $\eta(G)\leq 2k+1$. Let $e=uv$ be any edge in $G$. If
$N_k(u)\cap N_k(v)=\emptyset$, then $|V(G)|\geq
|N_k(u)|+|N_k(v)|+2>n$, a contradiction. Thus $N_k(u)\cap
N_k(v)\neq\emptyset$. Pick $w\in N_k(u)\cap N_k(v)$, and let $P$
(resp. $Q$) be a path between $u$ and $w$ (resp. between $v$ and
$w$). Then $e$ belongs a close trial $uPwQvu$ of length $2k+1$.
Therefore, $e$ belongs a cycle of length at most $2k+1$.
\end{proof}
\begin{remark} The above condition is almost optimal since $K_{n/2}\cup
K_{n/2}$ is disconnected for even $n$.
\end{remark}
\begin{corollary}
Let $G$ be a graph with minimum degree $\delta(G)$ and girth $g(G)$.
If there exists an integer $k$ such that $k\leq g(G)/2$ and
$\delta(G)(\delta(G)-1)^{k-1}>n/2-1$, then $G$ has an orientation
$H$ such that $rad(H)\leq 4k^2$.
\end{corollary}
\begin{proof}
Let $k$ be an integer such that $k\leq g(G)/2$ and
$\delta(G)(\delta(G)-1)^{k-1}>n/2-1$. For any vertex $u$ of $G$, let
$1\leq i<k$ be any integer and $x,y\in N_i(u)$. If $x$ and $y$ have
a common neighbor $z$ in $N_{i+1}(u)$, then $G$ has a cycle of
length at most $2i<2k\leq g(G)/2$, a contradiction. Thus $x$ and $y$
has no common neighbor in $N_{i+1}(u)$. Therefore, $|N_k(u)|\geq
\delta(G)(\delta(G)-1)^{k-1}>n/2-1$. By Theorem~\ref{T2}, $G$ has an
orientation $H$ such that $rad(H)\leq 4k^2$.
\end{proof}
\section{Upper bound for rainbow connection number}
At first, we have the following observation.
\begin{observation}
Let $G$ be a graph and $H$ be a spanning subgraph of $G$. Then
$rc(H)\leq rc(G)$.
\end{observation}
\noindent{\itshape Proof of Theorem~\ref{T4}:} Let $u$ be a center
of $G$ and let $H_0$ be the trivial graph with vertex set $\{u\}$.
We assert that {\itshape there exists a subgraph $H_i$ of $G$ such
that $N_i[u]\subseteq V(H_i)$ and $rc(H_i)\leq \Sigma_{j=1}^i
\min\{2(rad(G)-j)+1,\eta(G)\}$.}
{\itshape Basic step:} When $i=1$, we omit it since the proof of
this step is similar to that of the following induction step.
{\itshape Induction step:} Assume that the above assertion holds for
$i-1$ and $c$ is a\linebreak[3] $rc(H_{i-1})$-rainbow coloring of
$H_{i-1}$. Next we show that the above assertion holds for $i$. For
any $v\in N_i(u)$, either $v\in V(H_{i-1})$ or $v\in N(H_i)$ since
$N_{i-1}[u]\subseteq V(H_{i-1})$. If $N_i(u)\subseteq V(H_{i-1})$,
then let $H_i=H_{i-1}$ and we are done. Thus, we suppose
$N_i(u)\not\subseteq V(H_{i-1})$ in the following.
Let $C_1=\{\alpha_1,\alpha_2,\cdots\}$ and
$C_2=\{\beta_1,\beta_2,\cdots\}$ be two pools of colors, none of
which are used to color $H_{i-1}$. An edge-coloring of an $H$-ear
$P=(u_0,u_1,\cdots,u_k)$ is a $symmetrical\ coloring$ if its edges
are colored by $\alpha_1,\alpha_2,\cdots,\alpha_{\lceil
k/2\rceil},\beta_{\lfloor k/2\rfloor},\cdots,\beta_2,\beta_1$ in
that order or $\beta_1,\beta_2,\cdots,\beta_{\lfloor
k/2\rfloor},\alpha_{\lceil k/2\rceil}\cdots,\alpha_2,\alpha_1$ in
that order.
Let $X=N_i(u)\setminus V(H_{i-1})$ and
$m=\min\{2(rad(G)-i)+1,\eta(G)\}$. Pick $x_1\in X$, Let $y_1$ be a
neighbor of $x_1$ in $H_{i-1}$ and $P_1$ be an optimal
$(H_{i-1},x_1y_1)$-ear. We can color $P$ symmetrically with colors
$\alpha_1,\alpha_2,\cdots,\alpha_{\lceil
\ell(P)/2\rceil},\beta_{\lfloor
\ell(P)/2\rfloor},\ldots,\beta_2,\beta_1$. Pick $x_2\in X$
satisfying that all the incident edges of $x_2$ are not colored. Let
$y_2$ be a neighbor of $x_2$ in $H_{i-1}$. If there exists an
optimal $(H_{i-1},x_2y_2)$-ear $P_2$ such that $P_1$ and $P_2$ are
independent, then we can color $P_2$ symmetrically with colors
$\alpha_1,\alpha_2,\cdots,\alpha_{\lceil
\ell(P_2)/2\rceil},\beta_{\lfloor
\ell(P_2)/2\rfloor},\ldots,\beta_2,\beta_1$. Otherwise, by
Lemma~\ref{L1}, there exists an optimal $(H_{i-1},x_2y_2)$-ear
$P_2=P_{y_2z_2}$ such that $P_1$ and $P_2$ have only one continuous
common segment containing $z_2$, where $z_2$ is the other foot of
$P_2$. Thus we can color $P_2$ symmetrically with colors
$\alpha_1,\alpha_2,\cdots,\alpha_{\lceil
\ell(P_2)/2\rceil},\beta_{\lfloor
\ell(P_2)/2\rfloor},\ldots,\beta_2,\beta_1$ by preserving the
coloring of $P_1$. We can pick the vertices of $X$ and color optimal
$H_i$-ears until that for any $x\in X$, at least two incident edges
of $x$ are colored. Since for any leg $e$ of $H_{i-1}$, $\ell(e)\leq
m$ by Lemma~\ref{L3}, we use at most $m$ coloring in the above
coloring process.
Let $H_i$ be the graph obtained from $H_{i-1}$ by adding vertices in
$V(G)\setminus V(H_{i-1})$, which has at least two new colored
incident edges, and adding new colored edges. Clearly,
$N_i[u]\subseteq V(H_i)$. It is suffices to show that $H_i$ is
rainbow connected. Let $x$ and $y$ be two distinct vertices in
$H_i$. If $x,y\in V(H_{i-1})$, then there exists a rainbow path
between $x$ and $y$ by inductive hypothesis. If exactly one of $x$
and $y$ belongs to $V(H_{i-1})$, say $x$. Let $P$ be a symmetrical
colored $H_{i-1}$-ear containing $y$ and $y'$ be a foot of $P$.
There exists a rainbow path $Q$ between $x$ and $y'$ in $H_{i-1}$ by
inductive hypothesis. Thus, $xQy'Py$ is a rainbow path between $x$
and $y$ in $H_i$.
Suppose none of $x$ and $y$ belongs to $H_{i-1}$. Let $P$ and $Q$ be
symmetrical colored $H_{i-1}$-ear containing $x$ and $y$,
respectively. Furthermore, let $x',x''$ be the foot of $P$ and
$y',y''$ be the foot of $Q$. Without loss of generality, assume that
$P$ is colored from $x'$ to $x''$ by
$\alpha_1,\alpha_2,\cdots,\alpha_{\lceil\ell(P)/2\rceil},\beta_{\lfloor
\ell(P)/2\rfloor},\ldots,\beta_2,\beta_1$ in that order, and $Q$ is
colored from $y'$ to $y''$ by $\alpha_1,\alpha_2,
\cdots,\alpha_{\lceil \ell(Q)/2\rceil},\beta_{\lfloor
\ell(Q)/2\rfloor},\ldots,\beta_2,\beta_1$ in that order. If
$\ell(x'Px)\leq \ell(y'Qy)$. Let $R$ be a rainbow path between $x'$
and $y''$ in $H_{i-1}$. Then $xPx'Ry''Qy$ is a rainbow path between
$x$ and $y$ in $H_i$. Otherwise, $\ell(x'Px)> \ell(y'Qy)$. Let $R$
be a rainbow path between $y'$ and $x''$ in $H_{i-1}$. Then
$yPy'Rx''Qx$ is a rainbow path between $x$ and $y$ in $H_i$. Thus,
there exists a rainbow path between any two distinct vertices in
$H_i$, that is, $H_i$ is $(\Sigma_{j=1}^i
\min\{2(rad(G)-j)+1,\eta(G)\})$-rainbow connected. \hfill$\Box$
The following optimal example is from \cite{bas}.
\begin{figure}[h,t,b]
\begin{center}
\scalebox{0.8}[0.8]{\includegraphics{Fig3.eps}} \vspace{0.5cm}
Figure~3. Graph $H_{r,\eta(G)}$. Every $P_i$ is a path between $x_i$\\
and $x_{i-1}$ of length $\ell(P_i)=\min\{2i, \eta(G)-1\}$.
\end{center}
\end{figure}
\begin{example}\upshape
For any $r\geq 1$ and $3\leq\eta(G)\leq 2r + 1$, we first construct
the graph $H_{r,\eta(G)}$ as in Figure~3. Clearly, $H_{r,\eta(G)}$
is a bridgeless graph with radius $rad(G)=ecc(u)=r$ and every edge
of $H_{r,\eta(G)}$ is contained in a cycle of length at most
$\eta(G)$.
Let $m=\sum_{i=1}^r\min\{2i+1,\eta(G)\}$ and let $H^j$ be a copy of
$H_{r,\eta(G)}$, where $1\leq j\leq m^r+1$, and $V(H^j)=\{x^j : x\in
V(H_{r,\eta(G)})$ and $E(H_j)=\{x^jy^j\,|\, xy\in
E(H_{r,\eta(G)})\}$. Identify the vertex $u^j$ as a new vertex $u$.
The resulting graph is denoted by $G$. It is easy to check that $G$
is a bridgeless graph with radius $rad(G)=ecc(u)=r$ and every edge
of $H_{r,\eta(G)}$ is contained in a cycle of length at most
$\eta(G)$. Thus, $rc(G)\leq\Sigma_{i=1}^r \min\{2i+1, \eta(G)\}$ by
Theorem~\ref{T4}. On the other hand, for any $k< m$ and any $k$-edge
coloring of $G$, every $r$-length $P_{uv^j}$ path can be colored in
at most $k^r$ different ways. By the Pigeonhole Principle, there
exist $p\neq q,\,1\leq p<q\leq m^r+1$ such that
$c(x^p_{i-1}x^p_i)=c(x^q_{i-1}x^q_i)$ for $1\leq i\leq r$. Consider
any rainbow path $P$ from $v^p$ to $v^q$. For every $1\leq i\leq r$,
$x^p_{i-1}x^p_i$ belongs to $P$ if and only if $x^q_{i-1}x^q_i$ does
not belong to $P$. Thus, $\ell(P)\geq\Sigma_{i=1}^r \min\{2i+1,
\eta(G)\}=m$, and there does not exist any rainbow path between
$v^p$ and $v^q$. Hence, $rc(G)=\sum_{i=1}^r\min\{2i+1,\eta(G)\}$.
\end{example}
The following example shows that our result is better than that of
Theorem~\ref{T3}.
\begin{example}\upshape
Let $r\geq 3,k\geq 2r$ be two integers, and $W_k=C_k\vee K_1$ be an
wheel, where $V(C_k)=\{u_1,u_2,\ldots,u_k\}$ and $V(K_1)=\{u\}$. Let
$H$ be the graph obtained from $W_k$ by inserting $r-1$ vertices
between every edge $uu_i,\, 1\leq i\leq k$. For every edge $e=xy$ of
$H$, add a new vertex $v_e$ and new edges $v_ex,v_ey$. Denote by $G$
the resulting graph. It is easy to check that $rad(G)=r,\,
diam(G)=2r$, $\eta(G)=3$ and $\zeta(G)=2r-1$. By Theorem~\ref{T2},
we have $rc(G)\leq\sum_{i=1}^r\min\{2i+1,\zeta(G)\}\leq r^2+2r-2$.
But, by Theorem~\ref{T5} we have $rc(G)\leq 3r$. On the other hand,
$rc(G)\geq 2r$ since $diam(G)=2r$.
\end{example}
The remaining results are similar to those in Section~3.
\begin{theorem}
Let $G$ be a plane graph. If the length of the boundary of every
face is at most $k$, then $rc(G)\leq k\,rad(G)$.
\end{theorem}
\begin{corollary}
Let $G$ be a maximal plane (resp. outerplane) graph. Then $rc(G)\leq
3rad(G)$.
\end{corollary}
\begin{theorem}
Let $G$ be a bridgeless edge-transitive graph. Then $rc(G)\leq
rad(G)g(G)$, where $g(G)$ is the girth of $G$.
\end{theorem}
\begin{theorem}
Let $G=(V_1\cup V_2,E)$ be a bipartite graph with $|V_1|=n$ and
$|V_2|=m$. If $d(x)\geq k>\lceil m/2\rceil$ for any $x\in V_1$,
$d(y)\geq r>\lceil n/2\rceil$ for any $y\in V_2$, then $rc(G)\leq
12$.
\end{theorem}
\begin{remark}
The degree condition is optimal. Let $m, n$ be two even numbers with
$n, m\geq 2$. Since $K_{n/2,m/2}\cup K_{n/2,m/2}$ is disconnected,
$rc(K_{n/2,m/2}\cup K_{n/2,m/2})=\infty$.
\end{remark}
\begin{corollary}
Let $G=(V_1\cup V_2,E)$ be a $k$-regular bipartite graph with
$|V_1|=|V_2|=n$. If $k>\lceil n/2\rceil$, then $rc(G)\leq 12$.
\end{corollary}
The following theorem holds for general graphs.
\begin{theorem}
Let $G$ be a graph.
$(i)$ If there exists an integer $k\geq 2$ such that
$|N_k(u)|>n/2-1$ for every vertex $u$ in $G$, then $rc(G)\leq
4k^2+2k$.
$(ii)$ If $\delta(G)>n/2$, then $rc(G)\leq 6$.
\end{theorem}
\begin{remark}
The above condition is almost optimal since $K_{n/2}\cup K_{n/2}$ is
disconnected for even $n$.
\end{remark}
\begin{corollary}
Let $G$ be a graph with minimum degree $\delta(G)$ and girth $g(G)$.
If there exists an integer $k$ such that $k<g(G)/2$ and
$\delta(G)(\delta(G)-1)^{k-1}>n/2-1$, then then $rc(G)\leq 4k^2+2k$.
\end{corollary}
|
\section{Introduction}
A lot of interest, both in the mathematical and physical community,
has been devoted in the past 10 years to the study of the rotation of
a Bose Einstein condensate: experiments \cite{MCWD1,MCWD2},
theoretical works (see \cite{Coo,Fet} for reviews), mathematical
contributions \cite{Aft,LiSe}.
In the first experimental production of such rotating Bose Einstein
condensates, a rotating laser beam was superimposed on the magnetic
trap
holding the atoms in order to spin up the condensate by creating a
harmonic anisotropic rotating potential \cite{MCWD1,MCWD2}.
Recently, a new experimental device has emerged which consists in realizing artificial or synthetic gauge magnetic forces and leads to the formation of vortex lattices
at rest in the lab frame \cite{LCGPS}. A colloquium \cite{DGJO} has
analyzed in detail the artificial gauge fields and their
manifestations. In order to understand the main ingredients of the
physics of geometrical gauge fields, \cite{DGJO} (see also
\cite{GCYRD}) study the case of a single quantum particle
state with a 2 levels internal structure. More complex systems with
more than 2 internal levels are
also discussed in \cite{DGJO} but we stick here to the simpler case of 2 levels,
which contains all the mathematical difficulties.
A key issue is to determine whether one internal state can be followed adiabatically. A question raised by Jean Dalibard is to analyze in particular whether vortex formation may break down the adiabatic process. In \cite{DGJO}, some conditions are provided, that we want to analyze from a mathematical point of view.
We are interested in the minimization of the energy
\begin{eqnarray*}
\mathcal{E}_{\kappa}(\phi)&=&\int_{\field{R}^{2}}|\nabla \phi|^{2}+
V_{\kappa}(x,y)|\phi|^{2}+\frac{G}{2}|\phi|^{4}~dxdy\\
&&
\hspace{1cm}+
\Omega_{\kappa,\ell_{\kappa}}(x)\langle \phi\,,
\begin{pmatrix}
\cos(\theta_{\ell_{\kappa}}(x)) & e^{i\varphi_{k}(y)}\sin(\theta_{\ell_{\kappa}}(x))\\
e^{-i\varphi_{k}(y)}\sin(\theta_{\ell_{\kappa}}(x))&-\cos(\theta_{\ell_{\kappa}}(x))
\end{pmatrix}
\phi\rangle_{\field{C}} \newcommand{\nz}{\field{N}^{2}},
\end{eqnarray*}
where $\phi=
\begin{pmatrix}
\phi_{1}(x,y)\\\phi_{2}(x,y)
\end{pmatrix}
\in \field{C}} \newcommand{\nz}{\field{N}^{2}$ and
$|\phi(x,y)|^{2}=|\phi_{1}(x,y)|^{2}+|\phi_{2}(x,y)|^{2}$\,. We prescribe that the $L^2$ norm of $\phi$ is 1.
Here $\phi_1$ and $\phi_2$ are the internal degree of freedom of a particle: ground and excited state of the atom. It is assumed that the atom is shone by a quasi resonant laser beam.
The functions $\Omega_{\kappa,\ell_{\kappa}}(x)$, $\varphi_{k}(y)$ and $\theta_{\ell_{\kappa}}(x)$ are
given as in \cite{DGJO} by
\begin{eqnarray*}
&& \Omega_{\kappa,\ell_{\kappa}}(x)=\kappa\Omega(\frac{x}{\ell_{\kappa}})\hbox{ with }\Omega(x)=\sqrt{1+x^{2}}\,,
\\ && \theta_{\ell_{\kappa}}(x)=\underline{\theta}(\frac{x}{\ell_{\kappa}})\hbox{ with }
\cos(\underline{\theta}(x))=\frac{x}{\sqrt{x^{2}+1}} \quad,\quad
\sin(\underline{\theta}(x))=\frac{1}{\sqrt{x^{2}+1}}\,,\\
&& \varphi_{k}(y)=\underline{\varphi}(ky)\hbox{ with }
\underline{\varphi}(y)=y\,,
\end{eqnarray*} where $\underline\varphi $ is the phase of the propagating laser beam while $\underline\theta$ is the mixing angle.
We define \begin{eqnarray}\label{eq.defM}
M(x,y)=\Omega(x)
\begin{pmatrix}
\cos(\underline{\theta}(x)) & e^{i\underline{\varphi}(y)}\sin(\underline{\theta}(x))\\
e^{-i\underline{\varphi}(y)}\sin(\underline{\theta}(x))&-\cos(\underline{\theta}(x))
\end{pmatrix}\,.
\end{eqnarray}The matrix $M$ models the coupling between the atom and the laser.
The $2\times 2$ matrix
$M(\frac{x}{\sqrt{\ell_{\kappa}k}},\sqrt{\ell_{\kappa}k}y)$
can be diagonalized
in the bases $(\psi_{+},\psi_{-})$ respectively associated with the eigenvalues~$\pm \Omega(x)$,
\begin{equation}\label{psi+-}
\psi_{+}=
\begin{pmatrix}
C\\Se^{-i\varphi}
\end{pmatrix}
\;,\;
\psi_{-}=
\begin{pmatrix}
Se^{i\varphi}\\-C
\end{pmatrix}\,,
\end{equation}
with $C=\cos\left(\frac{1}{2}\,\underline{\theta}(\frac{x}{\sqrt{\ell_{\kappa}k}})\right)$,
$S=\sin\left(\frac{1}{2}\,\underline{\theta}(\frac{x}{\sqrt{\ell_{\kappa}k}})\right)$ and
$\varphi=\sqrt{\ell_{\kappa}k}\,y$. When the particle follows adiabatically the eigenstate $\psi_-$, this corresponds to set formally $\phi= u(x,y) \psi_-$, where
$ u(x,y)\in \field{C}} \newcommand{\nz}{\field{N}$ in $\mathcal{E}_{\kappa}$. Then $u$ minimizes a Gross-Pitaevskii type energy functional
with a modified trapping potential called the geometrical gauge potential.
The scalar
potential $V_{\kappa}(x,y)=V_{\kappa}(x,y)\Id_{\field{C}} \newcommand{\nz}{\field{N}^{2}}$ will be adjusted in order to produce a
harmonic potential after the addition of the geometrical gauge potential from the
adiabatic theory.
We want to justify the adiabatic approximation for states close to
$\psi_{-}$ and analyze the error term between the
initial and effective Hamiltonians.
After a rescaling, the parameter occurring in the experiments have the
following orders of magnitude
$$
\kappa\sim 10^{6}\quad,\quad G\sim 600\quad,
\quad\ell_{\kappa}\sim 25\quad,\quad k\sim 50\,,
$$
but other values can be discussed. Conditions on the strength and
spatial extent of the artificial potential have to be prescribed in
order to induce
large circulation. Two cases,
$\ell_{\kappa} k \geq 1$ and $\ell_{\kappa} k\leq 1$, can be
distinguished and the problem has to be rewritten in two different ways
in order to apply semiclassical techniques. In fact, we will focus on
the case $\ell_{\kappa}k\geq 1$, corresponding to the previous
numerical values.
The complete analysis is carried out in the asymptotic regime $\ell_{\kappa}k\to
+\infty$ but some partial results are also valid for
$\ell_{\kappa}k\leq 1$ or $\ell_{\kappa}k\to 0$\,.\\
A change of scale
$\phi(x,y)=\sqrt{\frac{k}{\ell_{\kappa}}}\psi(\sqrt{\frac{k}{\ell_{\kappa}}}x
, \sqrt{\frac{k}{\ell_{\kappa}}}y)$ yields a new expression for the energy:
\begin{multline*}
\int_{\field{R}^{2}} \frac{k}{\ell_{\kappa}}(|\partial_{x}\psi|^{2}+
|\partial_{y}\psi|^{2})+ V_{\kappa}(\sqrt{\frac{\ell_{\kappa}}{k}}x, \sqrt{\frac{\ell_{\kappa}}{k}}y)|\psi|^{2}
\\
+
\kappa\Omega(\frac{x}{\sqrt{k\ell_{\kappa}}})
\langle \psi\,,
\begin{pmatrix}
\cos(\underline{\theta}(\frac{x}{\sqrt{k\ell_{\kappa}}})) & e^{i\underline{\varphi}(\sqrt{k\ell_{\kappa}}y)}\sin(\underline{\theta}(\frac{x}{\sqrt{k\ell_{\kappa}}}))\\
e^{-i\underline{\varphi}(\sqrt{k\ell_{\kappa}}y)}\sin(\underline{\theta}(\frac{x}{\sqrt{k\ell_{\kappa}}}))&-\cos(\underline{\theta}(\frac{x}{\sqrt{k\ell_{\kappa}}}))
\end{pmatrix}\psi\rangle_{\field{C}} \newcommand{\nz}{\field{N}^{2}}\\
+ \frac{Gk}{2\ell_{\kappa}}|\psi|^{4}~dxdy\,.
\end{multline*}
According to the two cases $\ell_{\kappa} k\geq 1$ or $\ell_{\kappa} k\leq 1$, we
define a small parameter $\varepsilon$ that allows to rescale the energy. In fact, we define rather
the parameter $\varepsilon^{2+2\delta}$, where $\delta$ can be taken as a first step equal
to $5/2$\,. The exponent $\delta >0$ is a technical trick which
provides the right quantitative estimates for the adiabatic
approximation with a quadratic kinetic energy term\,. The suitable choice of this new parameter $\delta$ is
discussed further down in this introduction, in
Subsections~\ref{se.mainres} and~\ref{se.scheme}.
The small parameter $\varepsilon>0$ is thus introduced
according to the two cases:
\begin{description}
\item[if $\ell_{\kappa} k \geq 1$,] then
\begin{equation}
\label{eq.defeps1}
\varepsilon^{2+2\delta}= \frac{k^{2}}{\kappa}\;,\;
\delta >0\quad,\quad G_{\varepsilon}=\frac{Gk}{\kappa
\ell_{\kappa}}=\frac{G}{k\ell_{\kappa}}\varepsilon^{2+2\delta}\,;
\end{equation}
(This leads in our example to $\varepsilon^{2+2\delta}= 2.5~10^{-3}$.)
\item[if $\ell_{\kappa} k \leq 1$,] then
\begin{equation}
\label{eq.defeps2}
\varepsilon^{2+2\delta}= \frac{1}{\ell_{\kappa}^{2}\kappa}\;,\;
\delta >0\quad,\quad G_{\varepsilon}=\frac{Gk}{\kappa
\ell_{\kappa}}= Gk\ell_{\kappa}\varepsilon^{2+2\delta}.\end{equation}
\end{description} We define \begin{equation} \label{tau}\tau=(\tau_{x},\tau_{y})=
\left\{
\begin{array}[c]{ll}
(\frac{1}{k\ell_{\kappa}}, 1)&\text{if}~k\ell_{\kappa} \geq 1\,,\\
(1,k\ell_{\kappa})& \text{if}~k\ell_{\kappa} \leq 1\,.
\end{array}
\right.
\end{equation}
In both cases, this leads to \begin{equation}
\label{eq.rescEn1+}\kappa^{-1}{\mathcal E}_{\kappa}(\phi)=
{\mathcal E}_{\varepsilon}(\psi)
\end{equation} where
\begin{eqnarray}
\label{eq.defEeps}
&& \mathcal{E}_{\varepsilon}(\psi)=\langle \psi\,,\, H_{Lin}\psi\rangle
+ \frac{G_{\varepsilon,\tau}}{2}\int |\psi|^{4}~dxdy\,, \label{eq.a+L2}\\
&& G_{\varepsilon,\tau}=G\tau_{x}\tau_{y}\varepsilon^{2+2\delta}\,, \label{Geps}
\end{eqnarray}and \begin{eqnarray}
H_{Lin}&=&-\varepsilon^{2\delta}\tau_{x}\tau_{y}\varepsilon^{2}\Delta +
V_{\varepsilon,\tau}(x,y)+
M(\sqrt{\frac{\tau_{x}}{\tau_{y}}}x,\sqrt{\frac{\tau_{y}}{\tau_{x}}}y)
\label{Hlin}
\,,\\ \label{eq.defV}
V_{\varepsilon,\tau}(x,y)&=&\kappa^{-1}V_{\kappa}(\sqrt{\frac{\ell_{\kappa}}{k}}x,
\sqrt{\frac{\ell_{\kappa}}{k}}y)\quad (\textrm{to~be~fixed})\,.
\end{eqnarray}
The quadratic energy (linear Hamiltonian) is
\begin{eqnarray}
\nonumber
\mathcal{E}_{quad,
\varepsilon}(\psi)
&=&\int_{\field{R}^{2}}\varepsilon^{2+2\delta}\tau_{x}\tau_{y}|\nabla\psi|^{2}
+
V_{\varepsilon,\tau}(x,y)|\psi|^{2}
\\
\label{enerquad}
&&\hspace{1cm}+
\langle \psi\,,\, M(\sqrt{\frac{\tau_{x}}{\tau_{y}}}x,\sqrt{\frac{\tau_{y}}{\tau_{x}}}y)
\psi\rangle_{\field{C}} \newcommand{\nz}{\field{N}^{2}} ~dxdy
\\
&=& \langle \psi\,, H_{Lin} \psi\rangle\,.\label{Hlinener}
\end{eqnarray}
At least when $\tau_{x}=\tau_{y}=1$ and $\delta=0$, this problem looks like the standard problem of spatial adiabatic
approximation
studied in \cite{Sor}, \cite{MaSo}, \cite{MaSo2} \cite{PST}, although it requires some
adaptations because the symbols are neither bounded nor elliptic.\\
We shall consider the asymptotic analysis as $\varepsilon\to 0$ with
uniform control with respect to the parameters $G,k\ell_{\kappa}$ which
allow to fix the range of validity of the reduced models. Then we shall
consider the asymptotic behaviour of the reduced model and the whole
system as $\ell_{\kappa}k$ is large.
Specifying the right assumptions on $V_{\varepsilon,\tau}(x,y)$
or $V_{\kappa}$, possibly in a scale depending on $(\ell_{\kappa},k)$, is also an issue.
\subsection{Main result for the Gross-Pitaevskii energy}
\label{se.mainres}
We shall choose the potential $V_{\varepsilon,\tau}$ in (\ref{eq.defV}) such that after
the addition
of the adiabatic potential in the lower energy band, the effective potential is
almost harmonic. Namely, we assume
\begin{equation}
\label{eq.defVepsIntro}
V_{\varepsilon,\tau} (x,y)=
\frac{\varepsilon^{2+2\delta}}{\ell_{V}^{2}}v(\sqrt{\tau_{x}}x,\sqrt{\tau_{x}}y)
+
\sqrt{1+\tau_{x}x^{2}}-\varepsilon^{2+2\delta}
\left[
\frac{\tau_{x}^{2}}{(1+\tau_{x}x^{2})^{2}}+
\frac{1}{1+\tau_{x}x^{2}}
\right]\,,
\end{equation}
with the potential $v$ chosen such that \begin{eqnarray}
\label{eq.defvIntro}
&&v(x,y)=(x^{2}+y^{2})\chi_{v}(x^{2}+y^{2})+(1-\chi_{v}(x^{2}+y^{2}))\,
\\
\nonumber
&&
\chi_{v}\in \mathcal{C}^{\infty}_{0}([0,2))\quad, \quad 0\leq
\chi_{v}\leq 1\,, \quad \chi_{v}\equiv 1~\text{on}~[0,1],
\end{eqnarray}
and $\ell_V>0$ parametrizes the shape of the quadratic potential
around the origin.\\
With these assumptions on the potential $V_{\varepsilon,\tau}$,
if one chooses $\psi=u(x,y) \psi_-$, where $\psi_-$ is
the eigenfunction corresponding to $-\Omega (x\sqrt {\tau_x/\tau_y})$
of the matrix
$\Omega (x\sqrt {\tau_x/\tau_y})M (x\sqrt {\tau_x/\tau_y}, y\sqrt{\tau_y/\tau_x})$, then the linear part
$H_{Lin}$ in ${\cal E}_\varepsilon$ is formally replaced by the scalar $\varepsilon^{2+2\delta}\tau_x\hat H_-$
where \begin{equation}
\label{eq.defH-}
\hat{H}_{-}=
-\partial_{x}^{2}-
\left(\partial_{y}- i\frac{x}{2\sqrt{1+\tau_{x}x^{2}}}\right)^{2}
+
\frac{1}{\ell_{V}^{2}\tau_{x}}v(\sqrt{\tau_{x}}x,\sqrt{\tau_{x}}y)\,.
\end{equation} In the limit $\tau_x\to 0$, a natural $(\varepsilon,\tau)$-independent
scalar reduced model emerges:
$$
\mathcal{E}_{H}(u)=
\langle
u\,,\,
\left[-\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\right]u\rangle
+\frac{G}{2}\int_{\field{R}^2}|u|^{4}\,,\quad u(x,y)\in \field{C}} \newcommand{\nz}{\field{N}\,.
$$
Because $\kappa$ is large, or $\varepsilon$ is small, it is natural to expect that the ground state
of ${\cal E}_\varepsilon$ is close, up to a unitary transform, to a vector $u(x,y) \psi_-$. This is the aim of the adiabatic theory
and leads to a scalar problem. In order to get good bounds on the energy, we need to study the limit $\tau_x$ small
at the same time.
In all our work $\ell_{V}>0$ and $G>0$ are assumed to be fixed, while
the asymptotic behaviour is studied as $\varepsilon\to 0$ and
$\tau_{x}\to 0$\,.\\
The quadratic part of the above energy is associated with the
Hamiltonian
\begin{equation}
\label{eq.defHellV}
H_{\ell_{V}}= -\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\,, \quad (0<\ell_{V}<+\infty)\,,
\end{equation}
with the domain
\begin{equation}
{\mathcal H}_{2}=\left\{u\in L^{2}(\field{R}^{2})\,,
\sum_{|\alpha|+|\beta|\leq 2} \|q^{\alpha}D_{q}^{\beta}u\|_{L^{2}}
<+\infty\right\}\,,\quad q=(x,y)\in \field{R}^{2}\,,
\end{equation}
endowed with the norm
$\|f\|_{\mathcal{H}_{2}}^{2}=\sum_{|\alpha|+|\beta|\leq 2}
\|q^{\alpha}D_{q}^{\beta}u\|_{L^{2}}^{2}$ and the corresponding
distance $d_{\mathcal{H}_{2}}$\,.
It is not difficult (see Section~\ref{se.harmapp}) to check that the
minimization of $\mathcal{E}_{H}(\varphi)$ under the constraint
$\|\varphi\|_{L^{2}}=1$, admits solutions and that the set,
$\mathrm{Argmin}~\mathcal{E}_{H}$, of ground states for $\mathcal{E}_{H}$
is a bounded set of $\mathcal{H}_{2}$\,.
\begin{definition}
\label{de.min} For a functional $\mathcal{E}$ defined on a Hilbert
space $\mathcal{H}$ (with $+\infty$ as a possible value), we set
$$
\mathcal{E}_{min}=\inf_{\|u\|_{L^2}=1}\mathcal{E}(u)\quad
\text{and}\quad
\mathrm{Argmin}~\mathcal{E}=\left\{u\in \mathcal{H},\;
\mathcal{E}(u)=\mathcal{E}_{min}\,\text{and}\,\|u\|_{L^2}=1
\right\}\,.
$$
\end{definition}
\begin{theoreme}
\label{th.mintot}
Fix the constant $\ell_{V},G$ and $\delta$ and assume, for some $C_{\delta}=C(\ell_{V},G,\delta)>0$,
\begin{equation}
\label{eq.condtaueps}
\varepsilon^{2\delta}\leq \frac{\tau_{x}^{\frac{5}{3}}}{C_{\delta}}\,.
\end{equation}Let
$\chi=(\chi_{1},\chi_{2})$ be a pair of cut-off functions such that
$\chi_{1}^{2}+\chi_{2}^{2}\equiv 1$ on $\field{R}^{2}$, $\chi_{1}\in
\mathcal{C}^{\infty}_{0}(\field{R}^{2})$ and $\chi_{1}=1$ in a neighborhood
of $0$\,.\\
There exists $\nu_{0}=\nu_{\ell_{V},G}\in (0,\frac{1}{2}]$
and for any given $\delta>0$, there are constants
$\tau_{\delta}=\tau(\ell_{V},G,\delta)>0$\,,
$C_{\chi,\delta}=C(\ell_{V},G,\delta,\chi)>0$\,, and a unitary operator
$\hat{U}=\hat{U}(\varepsilon,\tau,\ell_{V},G,\delta)$ which
guarantee the following properties
\begin{itemize}
\item The energy $\mathcal{E}_{\varepsilon}$ introduced in
\eqref{eq.defEeps} admits ground states as soon as
$\tau_{x}\leq \tau_{\delta}$, with
$$
|\mathcal{E}_{\varepsilon,min}-\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{H,min}|\leq
C_{\delta}
\varepsilon^{2+2\delta}\tau_{x}^{\frac{5}{3}}\,.
$$
\item For any $\psi\in \mathrm{Argmin}~\mathcal{E}_{\varepsilon}$ written
in the form $\psi=\hat{U}
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$, the vector $a= \begin{pmatrix}a_{+}\\ a_-\end{pmatrix}$ satisfies
\begin{eqnarray*}
&&
\|a_{+}\|_{L^{2}}\leq C_{\delta}\varepsilon^{2+2\delta}\tau_{x}
\quad,\quad
\|a\|_{H^{2}}\leq \frac{C_{\delta}}{\tau_{x}}
\quad,\quad
\|a\|_{L^{4}}+\|a\|_{L^{6}}\leq C_{\delta}\,,
\\
&& \|\chi_{2}(\tau_{x}^{\frac{1}{9}}.)a_{-}\|_{L^{2}}\leq
C_{\chi,\delta}\tau_{x}^{\frac{1}{3}}\,,
\\
&&
d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.){a}_{-},
\mathrm{Argmin}~{\mathcal{E}_{H}})\leq
C_{\chi,\delta}(\tau_{x}^{\frac{2\nu_{0}}{3}}+\varepsilon)\,,\\
\\
&&
\|a_{+}\|_{L^{\infty}}\leq
C_{\delta}\varepsilon^{1+\delta}
\quad,\quad
d_{L^{\infty}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\mathrm{Argmin}~\mathcal{E}_{H})
\leq C_{\chi,\delta}(\tau_{x}^{\frac{2\nu_{0}}{3}}+\varepsilon)^{\frac{1}{2}}\,.
\end{eqnarray*}
\end{itemize}
All the constants can be chosen uniformly with respect to $\delta\in
(0,\delta_{0}]$ for any fixed $\delta_{0}>0$\,.
\end{theoreme}
\begin{remarque} The proof is made in two steps: 1) the limit
$\varepsilon\to 0$ corresponds to the adiabatic limit for the linear
problem and allows to replace the linear part $H_{Lin}$, of
$\mathcal{E}_{\varepsilon}$, by the scalar
$\varepsilon^{2+2\delta}\tau_{x}\hat{H}_{-}$ given by
\eqref{eq.defH-}\,;
2) the limit $\tau_x \to 0$ allows to reduce the asymptotic
minimization problem to a simpler one where the linear Hamiltonian
is exactly $H_{\ell_{V}}$ given by \eqref{eq.defHellV}.\\
One main point in the proof is to have
precise energy estimates for the limiting problem. In our case, we obtain them in the limit
$\tau_x\to 0$, because explicit calculations
are more easily accessible when the magnetic field is constant, that is
in the case of ${\mathcal E}_{H,min}$.
In theory, if one had precise energy estimates for a general $\tau_x$
for the intermediate adiabatic model with the linear part
$\hat{H}_{-}$,
complete results could be
performed for general values of~$\tau_x$\,.
\end{remarque}
\begin{remarque}
The exponent $\nu_{0}=\nu_{\ell_{V},G}$ is a Lojasiewicz-Simon
exponent (see Remark~\ref{re.loja} for an explanation and
a.e. \cite{Loj,BCR} for the definition of Lojasiewicz exponents and
\cite{Sim} for its extension to PDE problems). It is $\frac{1}{2}$ when the Lagrange multiplier
associated with $u\in \mathrm{Argmin}~\mathcal{E}_{H}$ is a simple
eigenvalue of $\hat{H}_{\ell_{V}}+G|\psi|^{2}$\,. There are reasons
to think that it is the case for generic values $(\ell_{V},G)\in
(0,+\infty)\times[0,+\infty)$, namely outside a subanalytic subset of dimension
smaller than $1$ (and possibly $0$). Nevertheless for $G=0$, there is
a discrete set of values of $\ell_{V}$ for which $H_{\ell_{V}}$ has multiple eigenvalues.
\end{remarque}
\noindent\textbf{Consequences and applications:}
One issue is whether the presence of vortices (zeroes of
the wave function with circulation around them) might break the
adiabatic approach. The answer contained in our theorem is that vortices of
$\psi$ in the original problem and vortices of the ground state of
the Gross-Pitaevskii energy ${\mathcal E}_H$ are close and that the
minimization of ${\mathcal E}_H$ provides all the information on the defects of
$\psi$. Indeed, the smallness of $a_+$ in $L^\infty$ indicates that the vortices
of $\psi$ and $a_-$ are close, and the last estimate of the theorem provides
that the vortices of $a_-$ are close to that of the Gross-Pitaevskii problem.
Numerically, in \cite{GCYRD},
the authors observe vortices in a system
with artificial gauge as presented in \cite{DGJO} and modeled with
the energy $\mathcal{E}_{\varepsilon}$. They check that the
vortex pattern is close to that of the Gross-Pitaevskii energy. If one wants to use
our results, one may process in the
following way: once $G$ is fixed, choose $\ell_{V}$ such that the
minimizers of $\mathcal{E}_{H}$ have vortices (detailed conditions
will be given in section \ref{se.harmapp}). Then
take $\tau_{x}>0$ and $\varepsilon>0$ small enough so that
the $L^{\infty}$ norm of $a_+$ and the $L^{\infty}$
distance from $\chi(\tau_{x}^{\frac{1}{9}}.)a_{-}$ to a ground state
of $\mathcal{E}_{H}$ are small.
In the above result, the constants are not explicitly controlled and this
control is worse and worse when $\delta$ increases. So it is not
explicit for given numerical values of the parameters
$\ell_{V},G,\tau_{x},\varepsilon$ or equivalently $\ell_{V},G,
\kappa,k,\ell_{\kappa}$\,. Nevertheless it provides a
framework for numerical
simulations, where the observation of vortices can be confirmed by
decreasing $\varepsilon$ and $\tau_{x}$\,. The parameters of the experiments
are just at the border where our constants may become too large
to provide a reasonable approximation.
The definitions \eqref{eq.defeps1} of $\varepsilon$ and
\eqref{tau} of $\tau_{x}$, transform the condition
\eqref{eq.condtaueps} into
$$
\left(\frac{k^{2}}{\kappa}\right)^{\frac{2\delta}{1+2\delta}}\ll \left(
\frac{1}{k\ell_{\kappa}}\right)^{\frac{5}{3}}\,.
$$
The larger $\delta$, the better, but $\delta$ must not be too
large because of the value of the constants $C_{\delta}$,
$C_{\delta,\chi}$. A value of a few units for $\delta$
does not affect them too much. Another reason for keeping $\delta$
small is that the initial small parameter is
$\varepsilon^{2+2\delta}$, while the final error estimate of
$d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\textrm{Argmin}~\mathcal{E}_{H})$ (or $d_{L^{\infty}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\textrm{Argmin}~\mathcal{E}_{H})$) contains
also an $\mathcal{O}(\varepsilon)$ term.\\
As an example for $\delta=\frac{5}{2}$, the above relation becomes
$$
\frac{1}{(k\ell_{\kappa})^{7/3}}\gg \frac{k^{2}}{\kappa} \quad,\quad
\tau_{x}=\frac{1}{k\ell_{\kappa}}
\quad,\quad \varepsilon=\left(\frac{k^{2}}{\kappa}\right)^{\frac{1}{7}}\,.
$$
A given precision of order $\varepsilon+\tau_{x}^{\frac{1}{3}}$, is
more easily achieved by taking $k$ small and
$\ell_{\kappa}=\frac{1}{\tau_{x}k}$ large (for example
$k=0.1$, $\ell_{\kappa}=500\times 25$ is better than the values $k=50$,
$\ell_{\kappa}=25$ given in the introduction). Note also that the
external potential $V_{\varepsilon,\tau}$, defined in
\eqref{eq.defVepsIntro}
must be adjusted up to the order $\varepsilon^{2+2\delta}=\frac{k^{2}}{\kappa}$\,.
\subsection{Gist of the analysis}
\label{se.scheme}
Following the general idea of
the founding articles \cite{BoFo,BoOp} of
Born, Fock and Oppenheimer, it is well known in the physics literature that a Hamiltonian system
$$
(\varepsilon D_{q})^{2} + u_{0}(q)
\begin{pmatrix}
E_{+}(q)&0\\
0&E_{-}(q)
\end{pmatrix}
u_{0}(q)^{*}
$$
is unitarily equivalent to a diagonal Hamiltonian plus a remainder term:
$$
\varepsilon^{2}
\left[
\begin{pmatrix}
(D_{q}- A(q))^{2}&0\\
0&(D_{q}+ A(q))^{2}
\end{pmatrix}
+|X(q)|^{2}\right] +R(\varepsilon)\,,
$$
where $(D_{q_{i}}\mp A_{i}(q))$ are the covariant derivatives ($\mp A(q)$ is the
adiabatic connection) associated with
the fiber bundles $u_{0}(q)
\begin{pmatrix}
\field{C}} \newcommand{\nz}{\field{N}\\0
\end{pmatrix}$ and $u_{0}(q)\begin{pmatrix}
0\\\field{C}} \newcommand{\nz}{\field{N}
\end{pmatrix}$, and $|X(q)|^{2}$ is the Born-Huang potential.\\
Since \cite{Kat} and until
recently (\cite{NeSo,MaSo,PST,PST2}), this
problem has been widely
studied by mathematicians,
and the remainder term
is formally:
$$
R(\varepsilon)=\sum_{i,j}\varepsilon^{2}C_{ij}(q)(\varepsilon
D_{q_{i}})(\varepsilon D_{q_{j}})+\mathcal{O}(\varepsilon^{3})\,.
$$
It is smaller than $\varepsilon^{2}$ for bounded frequencies (or
momenta) but it has the
same size as the main term
for a typical frequency of order $\frac{1}{\varepsilon}$\,.
For a nonlinear problem or without any information about the frequency
localization of the quantum states, it is important to estimate the
error terms in the low and high frequency regimes.\\
Introducing $\delta>0$ allows to obtain at the formal level
\begin{multline*}
\varepsilon^{2+2\delta}
\left[
\begin{pmatrix}
(D_{q}- A(q))^{2}&0\\
0&(D_{q}+ A(q))^{2}
\end{pmatrix}
+|X(q)|^{2}\right] +
\\
\sum_{i,j}\varepsilon^{2+4\delta}C_{ij}(q)(\varepsilon
D_{q_{i}})(\varepsilon D_{q_{j}})+\mathcal{O}(\varepsilon^{3+4\delta}) \,,
\end{multline*}
where the remainder term is now $\mathcal{O}(\varepsilon^{2+4\delta})$ at
the typical frequency $\frac{1}{\varepsilon}$ and thus
$\mathcal{O}(\varepsilon^{2\delta})$ times the size of the main term.
Such an error estimate can be made in the $L^{2}$-sense when applied
to some wave function $\psi$ lying in $\left\{|p|\leq
r\right\}$, with $p$ quantized into $\varepsilon D_{q}$, or more precisely fulfilling $\psi=\chi(\varepsilon
D_{q})\psi$ for some $\chi\in \mathcal{C}^{\infty}_{0}(\left\{|p|\leq
2r\right\})$ with
$$
\left\|R(\varepsilon) \psi\right\|\leq
C_{\chi}(\varepsilon^{2+4\delta}+
C_{N}\varepsilon^{N})\|\psi\|\leq (C'_{\chi,\delta}\varepsilon^{2\delta})\varepsilon^{2+2\delta}\|\psi\|\,.
$$
where $C_{N}$ essentially depends on the estimates of $N$-derivatives
of $u_{0}(q)$, $E_{\pm}(q)$\,. For a fixed $\delta$, we choose $N\geq
2+4\delta$\,.
\begin{remarque}
About the choice $\delta>0$,
another strategy could be considered in order to optimize the exponent
$\delta$, w.r.t $\varepsilon$:
under analyticity assumptions or more generally assumptions which lead
to an explicit control of $C_{N}$ in terms of $N$, one could think of
optimizing first
$C_{N}\varepsilon^{N}$ w.r.t to $N$ according to the methods of
\cite{NeSo,Sor,MaSo}. As an example with $C_{N}\leq N!$,
this would lead to
$C_{N}\varepsilon^{N}\leq Ce^{-\frac{1}{\varepsilon}}$ after choosing
$N=N(\varepsilon)=\left[\frac{1}{\varepsilon}\right]$, with some
$C\leq 1$\,.
Then taking
$\delta=\delta(\varepsilon)=-\frac{1}{4\varepsilon\log(\varepsilon)}$
would lead to
$$
CC_{\chi}(\varepsilon^{2+4\delta(\varepsilon)}+
e^{-\frac{1}{\varepsilon}})\leq 2CC_{\chi}e^{-\frac{1}{\varepsilon}}\|\psi\|\,.
$$
When $\delta(\varepsilon)\to \infty$ and as compared with
$\varepsilon^{2+2\delta(\varepsilon)}$ considered as the order $1$
term, an $\mathcal{O}(\varepsilon^{2+4\delta(\varepsilon)})$ remainder term
is almost of order $2$\,.\\
We do not consider this optimization of $\delta$ w.r.t $\varepsilon$,
with $\lim_{\varepsilon\to 0}\delta(\varepsilon)=+\infty$, because
our initial small parameter is $\varepsilon^{2+2\delta}$ and the
final result of Theorem~\ref{th.mintot}, (the estimate about $d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\textrm{Argmin}~\mathcal{E}_{H})$ in the nonlinear problem),
contains an $\mathcal{O}(\varepsilon)$\,. Hence we
keep $\delta>0$ independent of $\varepsilon$.
This is why $\delta$ is still present in the
constants of Theorem~\ref{th.mintot}.
\end{remarque}
We need to perform
a frequency (or momentum) truncation. We decompose a general $\psi$ into $\chi(\varepsilon
D_{q})\psi+(1-\chi(\varepsilon D_{q}))\psi$ and use rough estimates
for the part $(1-\chi(\varepsilon D_{q}))\psi$ which will be
compensated in the minimization problem for
$\mathcal{E}_{\varepsilon}$ by good a priori estimates for the norm
$\|(1-\chi(\varepsilon D_{q}))\psi\|$ of the
high-frequency part.
We also have to check that the unitary transform
$\hat{U}$, implementing the adiabatic approximation, does not perturb too much the nonlinear part of
$\mathcal{E}_{\varepsilon}(\psi)$\,.
In our case, the limit $\varepsilon\to 0$ leads to the Born-Oppenheimer Hamiltonians
$$
-\partial_{x}^{2}-(\partial_{y}\mp\frac{ix}{2\sqrt{1+\tau_{x}x^{2}}})^{2}
+ \frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}\tau_{x}}\,,
$$
which, in a second step, in the limit $\tau_{x}\to 0$,
leads to $H_{\ell_{V}}$ (with the sign $-ix$ for the lower energy band). This means that the
convergence to the Born-Oppenheimer Hamiltonian as
$\varepsilon\to 0$ has to be uniform w.r.t $\tau_{x}\in (0,1]$\,. This
last point requires to reconsider carefully the work of \cite{PST} by
following the uniformity w.r.t $\tau$ of the estimates given by
Weyl-H{\"o}rmander calculus (\cite{Hor,BoLe}) for $\tau$-dependent
metrics which have uniform structural constants.
\\
This is done for the low frequency part in Section~\ref{se.BO} while the basic tools of
semiclassical calculus are reviewed and adapted in the Appendix~\ref{se.semiclass}.\\
In Section~\ref{se.minipb} the error associated in the high-frequency
part is considered, as well as the effect of the unitary adiabatic
transformation on the nonlinear term.\\
Once the adiabatic approximation is well justified in this rather
involved framework, the accurate analysis, as well as the comparison when
$\tau_{x}$ is small, of the two reduced models (the one with
$H_{\ell_{V}}$ and the one with $\hat{H}_{-}$) is carried out in
Section~\ref{se.explnlmin-}. This follows the general scheme of
comparison of minimization problems: 1) Write energy estimates;
2) Use bootstrap arguments and possibly Lojasiewicz-Simon inequalities
in order to compare the minimizers in the energy space; 3) Use
the Euler-Lagrange equations in order to get a better comparison in
higher regularity spaces.\\
In Section~\ref{se.compmini}, all the information of the previous
sections is gathered in order to prove Theorem~\ref{th.mintot} :
existence of a minimizer for ${\mathcal E}_\varepsilon$ in Proposition \ref{pr.existEeps}, key energy estimates in Proposition~\ref{pr.compen}
and bounds for minimizers in Proposition~\ref{pr.adEL}.\\
Some comments and additional results are pointed out in
Section~\ref{se.complements}, namely: 1) the question of the smallness
condition of $\varepsilon$ w.r.t $\tau_{x}$ appearing in
Theorem~\ref{th.mintot}; 2) the possible extension to anisotropic
nonlinearities (one would have to check that the unitary transform
implementing the adiabatic approximation does not perturb the nonlinear part); 3) the minimization problem for excited states,
i.e. locally and approximately carried by $\psi_{+}$ instead of $\psi_{-}$; 4) the
extension to the problem of the time nonlinear dynamics of
adiabatically prepared states.
\section{Adiabatic approximation for the linear problem}
\label{se.BO}
In \cite{PST}, the adiabatic approximation is completely justified for
bounded symbols or when global elliptic properties of the complete
matricial symbol allow to reduce to this case after spectral
truncation. Unfortunately, it is not
the case here, because the eigenvalues of the symbol of the linear part
$H_{Lin}$ are $\varepsilon^{2\delta}\tau_{x}\tau_{y}|p|^{2}+ V_{\varepsilon,\tau}(x,y)\pm
\Omega(\sqrt{\frac{\tau_{x}}{\tau_{y}}}x)$\,. In \cite{Sor}, the adiabatic theory for unbounded symbols
is developed after stopping the complete asymptotic expansion in an
optimal way, under some analyticity assumptions, but this would be
particularly tricky here with divergences occurring both in the momentum
and position directions. We shall see that the sublinear divergence in
position makes no difficulty after using the right Weyl-H{\"o}rmander
class. The quadratic divergence in momentum, with the kinetic energy
$\tau_{x}\tau_{y}|p|^{2}$ is solved by first considering truncated kinetic energies and
using the additional scaling factor $\varepsilon^{2\delta}$,
$\delta>0$, in front of the kinetic energy term.\\
Our problem shows an anisotropy in the position variables $(x,y)$\,. The analysis of the
linear problem can be treated in $\field{R}^{d}$\,. Then we split the
position and momentum variables, $q\in\field{R}^{d}$ and $p\in\field{R}^{d}$,
into:
$$
q=(q',q'')\quad,\quad p=(p',p'')\quad q',p'\in \field{R}^{d'}\,,\,
q'',p''\in\field{R}^{d''}\quad,\quad d'+d''=d\,,
$$
and the pair $(\tau_{x},\tau_{y})$ is accordingly denoted by
$(\tau',\tau'')\in (0,1]^{2}$\,.\\
From this section, some notions and notations related with
semiclassical analysis are used. In particular, the notation
$S_{u}(m_{\tau},g_{\tau})$ refers to classes of
$(\varepsilon,\tau)$-dependent
symbols of which the seminorms are \underline{u}niformly controlled
w.r.t to the parameters $(\varepsilon,\tau)\in
(0,\varepsilon_{0})\times(0,1]^{2}$\,. For accurate definitions, we refer the reader to
appendix~\ref{se.semiclass} where all the necessary material is
reviewed and adapted for our analysis, assuming knowledges about
Fr{\'e}chet spaces and generalized functions.
\subsection{Born-Oppenheimer Hamiltonian}
Consider the Hamiltonian in $\hat{H}_{\varepsilon}=H(q,\varepsilon
D_{q},\varepsilon)$
with the symbol on $\field{R}^{2d}_{q,p}$
\begin{eqnarray*}
&&H(q,p,\varepsilon)=\varepsilon^{2\delta}\tau'\tau''|p|^{2}\gamma(\tau'\tau''|p|^{2})+\mathcal{V}(q,\tau,\varepsilon)
\\
&&\;=
\varepsilon^{2\delta}\tau'\tau''|p|^{2}\gamma(\tau'\tau''|p|^{2})
+u_{0}(q,\tau,\varepsilon)
\begin{pmatrix}
E_{+}(q,\tau,\varepsilon)& 0\\
0 & E_{-}(q,\tau,\varepsilon)
\end{pmatrix}
u_{0}^{*}(q,\tau,\varepsilon)\,,
\end{eqnarray*}
with $\delta>0$ and $(\tau',\tau'')\in (0,1]^{2}$\,.
The following properties, with the splitting of variables
$q=(q',q'')\in \field{R}^{d}$, are assumed:
\begin{eqnarray}
\nonumber
&&
E_{\pm}\in S_{u}\Big(\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle,
\frac{\frac{\tau'}{\tau''}d{q'}^{2}}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle^{2}}+
\frac{\tau''}{\tau'}d{q''}^{2}\Big)
\;,\\
&&
\label{eq.hypellEpm}
E_{+}(q,\tau,\varepsilon)-E_{-}(q,\tau\varepsilon)\geq C^{-1}\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle\,,\\
&&
\nonumber
u_{0}=(u_{0}^{*})^{-1}\in S_{u}(1,
\frac{\frac{\tau'}{\tau''}d{q'}^{2}}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle^{2}}+
\frac{\tau''}{\tau'}d{q''}^{2}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\\
&&
\nonumber
\quad \gamma\in \mathcal{C}^{\infty}_{0}(\field{R};\field{R}_{+})\quad\text{with}\quad
\gamma\equiv 1 \in [0,2r_{\gamma}^{2}]\,.
\end{eqnarray}
With these assumptions we are able to justify the Born-Oppenheimer
adiabatic approximation for $\delta>0$, $\tau',\tau''\in (0,1]$.
We shall work with the $\tau$-dependent metric
$$
g_{\tau}=
\frac{\frac{\tau'}{\tau''}d{q'}^{2}}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle^{2}}+
\frac{\tau''}{\tau'}d{q''}^{2}
+\frac{\tau'\tau''dp^{2}}{\langle
\sqrt{\tau'\tau''} p\rangle^{2}}\quad,
$$
on the phase-space $\field{R}^{2d}_{q,p}$, which is checked to have uniform
properties w.r.t $\tau\in (0,1]^{2}$ in Proposition~\ref{pr.metadm}\,.
The exact definition of the
parameter dependent H{\"o}rmander symbol classes,
$S_{u}(m,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, is given in
Appendix~\ref{se.semiclass}
(see in particular its meaning for $\tau$-dependent metrics in
Appendix~\ref{se.variametric}). We shall use the notation
$$
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle \langle
\sqrt{\tau'\tau''} p
\rangle^{\infty}},
g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)
$$ with the meaning of the exponent $\infty$ being
the same as in
$\mathcal{C}^{\infty}$ of $\mathcal{O}(\varepsilon^{\infty})$\,.
\begin{theoreme}
\label{th.BornOp}
There exists a unitary operator
$\hat{U}=U(q,\varepsilon D_{q},\tau, \varepsilon)$ with
symbol
\begin{eqnarray*}
&&U(q,p,\tau,\varepsilon)=u_{0}(q,\tau,\varepsilon) + \varepsilon
u_{1}(q,p,\tau,\varepsilon)+ \varepsilon^{2}u_{2}(q,p,\tau,\varepsilon)
\,,\\
&&\varepsilon^{-2\delta}u_{1,2}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle \langle
\sqrt{\tau'\tau''} p
\rangle^{\infty}},
g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,,
\end{eqnarray*}
such that
\begin{multline}
\label{eq.BornOpp1}
\hat{U}^{*}\hat{H}\hat{U}
=
\begin{pmatrix}
h_{BO,+}(q,\varepsilon D_{q},\varepsilon)&0\\
0& h_{BO,-}(q,\varepsilon D_{q},\varepsilon)
\end{pmatrix}\\
+ \varepsilon^{2+4\delta}R_{1}(q,\varepsilon
D_{q},\tau,\varepsilon)+ \varepsilon^{3+2\delta}R_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)\,,
\end{multline}
with $h_{BO\pm}(q,p,\tau,\varepsilon)$ equal to
\begin{eqnarray}
\nonumber
&=&\varepsilon^{2\delta}\tau'\tau''(|p
\mp \varepsilon
A|^{2}+|\varepsilon X|^{2}) + E_{\pm}
\\
\label{eq.hBOpm}
&=&\varepsilon^{2\delta}\tau'\tau''\left[\sum_{k=1}^{d}(p_{k}\mp \varepsilon A_{k})^{2}
+|\varepsilon X_{k}|^{2}\right]
+ E_{\pm}\,,
\quad
\end{eqnarray}
when $\sqrt{\tau'\tau''}|p|\leq r_{\gamma}$, and
$$
\begin{pmatrix}
+A_{k}&X_{k}\\
\bar{X_{k}}&-A_{k}
\end{pmatrix}=iu_{0}^{*}(\partial_{q^{k}}u_{0})\,.
$$
The remainder terms satisfy $R_{1},R_{2}\in
S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ and $R_{2}$ vanishes in
$\left\{\sqrt{\tau'\tau''}|p|\leq r_{\gamma}\right\}$ and those
estimates are uniform w.r.t $\delta\in
(0,\delta_{0}]$ (and $\tau\in (0,1]^{2}$, $\varepsilon\in (0,\varepsilon_{0}]$).
\end{theoreme}
\begin{remarque}
\begin{itemize}
\item
The remainder term is really negligible only for
$\delta>0$. This is explained in Subsection~\ref{se.discuss}.
\item The (Weyl)-quantization of $\varepsilon^{2\delta}\tau'\tau''(|p\mp \varepsilon
A|^{2}+|\varepsilon X|^{2})+E_{\pm}$ is nothing but
$$
\varepsilon^{2+2\delta}\tau'\tau''\left[\sum_{k=1}^{d}-(\partial_{q^{k}}\mp
iA_{k})^{2}+|X_{k}|^{2}\right]
+ E_{\pm}\,.
$$
\end{itemize}
\end{remarque}
\subsection{Second order computations for space adiabatic approximate projections of
the reduced Hamiltonian}
\label{se.secondord}
We shall consider the matricial symbol, on $\field{R}^{2d}_{q,p}$,
$$
H(q,p,\tau,\varepsilon)=
f_{\varepsilon}(p,\tau)
+ E_{+}(q,\tau,\varepsilon)\Pi_{0}(q,\tau,\varepsilon)
+ E_{-}(q,\tau,\varepsilon)(1- \Pi_{0}(q,\tau,\varepsilon))
$$
where $\Pi_{0}(q,\tau,\varepsilon)=\Pi_{0}(q,\tau,\varepsilon)^{*}=\Pi_{0}(q,\tau,\varepsilon)^{2}\in {\mathcal
M}_{2}(\field{C}} \newcommand{\nz}{\field{N})$\,,
$\gamma,E_{\pm}$ real-valued,
with $\delta \geq 0$ and the following properties:
\begin{eqnarray}
&&
\label{eq.hypEPi1}
E_{\pm}\in S_{u}(\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle,
g_{q\tau})
\quad,\quad
\Pi_{0}\in S_{u}(1, g_{q,\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\\
\nonumber
\text{with}&&
g_{q,\tau}=\frac{\frac{\tau'}{\tau''}d{q'}^{2}}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle^{2}}+
\frac{\tau''}{\tau'}d{q''}^{2},
\\
&&
\label{eq.hypEPi2}
E_{+}(q,\tau,\varepsilon)-E_{-}(q,\tau,\varepsilon)\geq C^{-1}\langle
\sqrt{\frac{\tau'}{\tau''}}
q'\rangle\,,\\
&&
\nonumber
f_{\varepsilon}(p,\tau)=\varepsilon^{2\delta}f_{1}(\sqrt{\tau'\tau''}p)\quad,
\quad f_{1}\in \mathcal{C}^{\infty}_{0}(\field{R}^{d};\field{R})
\quad,\quad \delta \geq 0\,.
\end{eqnarray}
For conciseness, the arguments $p$, $q$ and the parameters
$\tau,\varepsilon$,
will often be omitted in
$\partial_{p}^{\alpha}f_{\varepsilon}(p)$ and
$\partial^{\alpha}_{q}E_{\pm}(q)$ or $\partial^{\alpha}_{q}\Pi_{0}$\,.\\
According to Appendix~\ref{se.variametric}, the metric
$$
g_{\tau}=g_{q,\tau}
+ \frac{\tau'\tau''dp^{2}}{\langle
\sqrt{\tau'\tau''}p\rangle^{2}}
=
\frac{\frac{\tau'}{\tau''}d{q'}^{2}}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle^{2}}
+
\frac{\tau''}{\tau'}d{q''}^{2}
+
\frac{\tau'\tau''dp^{2}}{\langle
\sqrt{\tau'\tau''}p\rangle^{2}}
$$
has the gain function
$$
\lambda(q,p)=\min\left(\frac{\langle \sqrt{\tau'\tau''}p\rangle\langle
\sqrt{\frac{\tau'}{\tau''}}q\rangle}{\tau'},\frac{\langle
\sqrt{\tau'\tau''}p\rangle}
{\tau''}\right)\geq \langle \sqrt{\tau'\tau''}p\rangle\,.
$$
The $\varepsilon$-quantized version of the symbol $A_{\varepsilon}(q,p)$ will be denoted
$$
\hat A= A(q,\varepsilon D_{q})\,.
$$
Note that the symbols $E_{+}-E_{-}$ is elliptic in
its class $S_{u}(\langle \sqrt{\frac{\tau'}{\tau''}}q'
\rangle, g_{\tau})$\,,
and
$$
\varepsilon^{-2\delta}f_{\varepsilon}\in
S_{u}(\frac{1}{\langle \sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau})\,.
$$
Our aim is to compute accurately the adiabatic projection
$\Pi^{(n)}(q,p)=\Pi_{0}(q,\varepsilon)+\varepsilon \Pi_{1}(q,p,\varepsilon)+\cdots +
\varepsilon^{n}\Pi_{n}(q,p,\varepsilon)$ such that
$$
\hat \Pi^{(n)}\circ \hat\Pi^{(n)}= \hat \Pi^{(n)}+{\mathcal O}(\varepsilon^{n+1})
\quad,\quad
\left[
\hat H, \hat \Pi^{(n)}\right]= {\mathcal O}(\varepsilon^{n+1})\,.
$$
The general theory presented in \cite{PST}, tells us
that the asymptotic expansion can be
pushed up to $n=\infty$, but we will do here
accurate calculations up to
$n=2$ (with additional information for $n=3$) and then discuss the influence of the factor
$\varepsilon^{2\delta}$\,.
Those are feasible and rather easy because the kinetic
energy term and the two-level potential are simple.
This allows to
reconsider accurately the arguments sketched in \cite{PST} for the
Born-Oppenheimer
case with all the technical new peculiarities of our example. Note
also that in \cite{MaSo2}~chapter~10 explicit calculations have been made up to
order $n=5$ but in the slightly different framework of time-dependent
Born-Oppenheimer approximation oriented to polyatomic molecules~:~no use of the
exponent $\delta>0$, no divergence as $q\to \infty$ and no ellipticity
problem, no
extra-parameter $\tau$ and the techniques
are slightly different although still relying on semiclassical calculus.\\
Like in \cite{PST}, \cite{Sor}, \cite{MaSo}, \cite{MaSo2}, the calculations are first
done at the symbolic level and we write
$$
\left(A(\varepsilon)=\mathcal{O}_{S}(\varepsilon^{\nu})\right)
\Leftrightarrow
\left(
\varepsilon^{-\nu}A\in
S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))
\right)\,.
$$
For a matricial symbol $A(q,p)$ (possibly depending on $(\varepsilon,\tau)$), it is convenient to introduce the
diagonal and off-diagonal parts
\begin{eqnarray}
\label{eq.diag}
A^{D}(q,p)&:=&\Pi_{0}(q)A(q,p)\Pi_{0}(q)+ (1-\Pi_{0}(q))A(q,p)(1-\Pi_{0}(q))
\\
\label{eq.offdiag}
A^{OD}(q,p)&:=&\Pi_{0}(q)A(q,p)(1-\Pi_{0}(q))+(1-\Pi_{0}(q))
A(q,p)\Pi_{0}(q)\,,
\end{eqnarray}
where we recall $\Pi_{0}(q)=\Pi_{0}(q,\tau,\varepsilon)$\,.
Note the equalities
\begin{equation}
\label{eq.diagprod}
(AB)^{D}=A^{D}B^{D}+A^{OD}B^{OD}\quad, \quad (AB)^{OD}=A^{D}B^{OD}+A^{OD}B^{D}\,.
\end{equation}
The ``Pauli matrix''
\begin{equation}
\label{eq.sigz}
\sigma_{3}(q,\tau,\varepsilon)=2\Pi_{0}(q,\tau,\varepsilon)-\Id_{\field{C}} \newcommand{\nz}{\field{N}^{2}}\,,
\end{equation}
will also be used,
with the relations
\begin{eqnarray*}
&&\sigma_{3}^{2}(q)=\Id_{\field{C}} \newcommand{\nz}{\field{N}^{2}}\,,\quad
\sigma_{3}(q) A^{D}(q,p)\sigma_{3}(q)= A^{D}(q,p)\,,\\
&&\sigma_{3}(q)A^{OD}(q,p)\sigma_{3}(q)=-A^{OD}(q,p)\,,\\
&&
\sigma_{3}(q)A^{OD}(q,p)=\Pi_{0}(q)A(q,p)(1-\Pi_{0}(q))-(1-\Pi_{0}(q))A(q,p)\Pi_{0}(q)\,.
\end{eqnarray*}
We are looking for
\begin{eqnarray*}
&&\Pi^{(n)}(q,p,\tau,\varepsilon)
=
\sum_{j=0}^{n}\varepsilon^{j}\Pi_{j}(q,p,\tau,\varepsilon)\quad\in S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\\
\text{with}
&&
\Pi_{j}\in
S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\end{eqnarray*}
and such that
\begin{eqnarray}
\label{eq.proj}
&& \Pi^{(n)}\sharp^{\varepsilon}\Pi^{(n)}-\Pi^{(n)}=
\mathcal{O}_{S}(\varepsilon^{n+1})\,,\\
\label{eq.adj}
&&
{\Pi^{(n)}}^{*}=\Pi^{(n)}\,,
\\
\label{eq.comm}
&&
H\sharp^{\varepsilon}\Pi^{(n)}-\Pi^{(n)}\sharp H=\mathcal{O}_{S}(\varepsilon^{n+1})
\,.\end{eqnarray}
Like in \cite{MaSo}, \cite{PST}, this system is solved by induction by
starting from
$\Pi^{(0)}(q,p,\tau,\varepsilon)=\Pi_{0}(q,\tau,\varepsilon)$, with
\begin{eqnarray}
\label{eq.defGn1}
G_{n+1}&:=&
\varepsilon^{-(n+1)}\left[\Pi^{(n)}\sharp^{\varepsilon}\Pi^{(n)}-\Pi^{(n)}\right]
\mod
\mathcal{O}_{S}(\varepsilon)\,,\\
\label{eq.Pin1D}
\Pi_{n+1}^{D}&:=&-\sigma_{3}G_{n+1}^{D}\,,\\
\nonumber
F_{n+1}&:=&\varepsilon^{-(n+1)}
\left[H\sharp^{\varepsilon}(\Pi^{(n)}+\varepsilon^{n+1}\Pi_{n+1}^{D})\right.\\
\nonumber
&&
\hspace{3cm}
\left.-(\Pi^{(n)}+\varepsilon^{n+1}\Pi_{n+1}^{D})\sharp^{\varepsilon}H\right]\mod\mathcal{O}_{S}(\varepsilon)
\\
\label{eq.defFn1}
&=&
\varepsilon^{-(n+1)}
\left[H\sharp^{\varepsilon}\Pi^{(n)}-\Pi^{(n)}\sharp^{\varepsilon}H\right]
\mod\mathcal{O}_{S}(\varepsilon)\,,
\\
\label{eq.Pin1OD}
\Pi_{n+1}^{OD}&:=&
-\frac{1}{E_{+}(q)-E_{-}(q)} \sigma_{3}F_{n+1}^{OD}=\frac{1}{E_{+}(q)-E_{-}(q)} F_{n+1}^{OD}\sigma_{3}\,.
\end{eqnarray}
The general theory says that the principal symbol of
$F_{n+1}$ is off-diagonal,
$F_{n+1}=F_{n+1}^{OD}\mod\mathcal{O}_{S}(\varepsilon)$\,,
and $F_{n+1}$ can be chosen so that
\begin{equation}
\label{eq.FOD}
F_{n+1}=F_{n+1}^{OD}\,.
\end{equation}
Below are the computations up to $n=2$ in our specific case. In these
computations, we shall use Einstein's summation rule
$s_{k}t^{k}=\sum_{k}s_{k}t^{k}$ with
the coordinates $(p_{k},q^{k})$ or $(p^{k},q^{k})$ with
$p^{k}=p^{\ell}\delta_{\ell,k}=p_{k}$ like in the examples
$$
|p|^{2}=p^{k}p_{k}=p_{k}p_{\ell}\delta^{\ell,k}=p^{k}p^{\ell}\delta_{k,\ell}\quad,\quad
(\partial_{p}f_{\varepsilon}).\partial_{q}=(\partial_{p^{k}}f_{\varepsilon})\delta^{k,\ell}\frac{\partial}{\partial_{q^{\ell}}}=(\partial_{p_{k}}f_{\varepsilon})\partial_{q^{k}}\,.
$$
\noindent{$\mathbf{n=0}$:} Start with $\Pi^{(0)}=\Pi_{0}(q,\tau,\varepsilon)$ and notice
$\partial_{p}\Pi_{0}\equiv 0$ and $\partial_{q}\Pi_{0}\equiv (\partial_{q}\Pi_{0})^{OD}$\,.\\
\noindent{$\mathbf{n=1}$:} Take
\begin{eqnarray*}
&&G_{1}=\Pi_{0}\circ \Pi_{0}-\Pi_{0}=0\,,
\\
\text{and}&&
\Pi_{1}^{D}=0\,.
\end{eqnarray*}
Next compute
\begin{eqnarray*}
\varepsilon^{-1}\left[H\sharp^{\varepsilon}\Pi_{0}(q,\varepsilon)-\Pi_{0}(q,\varepsilon)\sharp^{\varepsilon}H\right]
&=&\varepsilon^{-1}\left[f_{\varepsilon}\sharp^{\varepsilon}\Pi_{0}(q,\varepsilon)-\Pi_{0}(q,\varepsilon)
\sharp^{\varepsilon}f_{\varepsilon}\right]\\
&=& -i\partial_{p_{k}}f_{\varepsilon}\partial_{q^{k}}\Pi_{0}
\mod \mathcal{O}_{S}(\varepsilon)\,,
\end{eqnarray*}
and take
\begin{eqnarray*}
&&
F_{1}=-i \partial_{p_{k}}f_{\varepsilon}\partial_{q^{k}}\Pi_{0}=
-i\partial_{p_{k}}f_{\varepsilon}(\partial_{q^{k}}\Pi_{0})^{OD}\,,
\\
\text{and}&&
\Pi^{(1)}=\Pi_{0}+
\frac{\varepsilon i\partial_{p_{k}}f_{\varepsilon}}{E_{+}-E_{-}}
\sigma_{3}(\partial_{q^{k}}\Pi_{0})^{OD}\,,
\end{eqnarray*}
with
$$
\varepsilon^{-2\delta}\Pi_{1}=\frac{i\partial_{p_{k}}f_{1}}{E_{+}-E_{-}}
\sigma_{3}(\partial_{q^{k}}\Pi_{0})^{OD}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle
\langle \sqrt{\tau'\tau''}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,.
$$
\noindent$\mathbf{n=2}$: Consider now
\begin{eqnarray*}
\varepsilon^{2}G_{2}&=&\Pi^{(1)}\sharp^{\varepsilon}
\Pi^{(1)}-\Pi^{(1)}\mod\mathcal{O}_{S}(\varepsilon^{3})\\
&=&\Pi_{0}\sharp^{\varepsilon} \Pi_{0}-\Pi_{0}+\varepsilon(\Pi_{0}\sharp^{\varepsilon}
\Pi_{1}+\Pi_{1}\sharp^{\varepsilon} \Pi_{0})-\Pi_{1}+\varepsilon^{2}\Pi_{1}\sharp^{\varepsilon}\Pi_{1}
\mod\mathcal{O}_{S}(\varepsilon^{3})
\,.
\end{eqnarray*}
According
to \eqref{eq.Wexpanexp}, with
$\Pi_{0}\sharp \Pi_{0}=\Pi_{0}$ and with
$$
\Pi_{0}\Pi_{1}+\Pi_{1}\Pi_{0}=
\Pi_{0}^{D}\Pi_{1}^{OD}+\Pi_{1}^{OD}\Pi_{0}^{D}
=
\Pi_{1}\,,
$$
we
can take
$$
G_{2}=
\frac{1}{2i}\left[-\partial_{q^{k}}\Pi_{0}\partial_{p_{k}}\Pi_{1}+\partial_{p_{k}}\Pi_{1}\partial_{q^{k}}\Pi_{0}\right]
+ \Pi_{1}^{2}\,.
$$
The first term $-\partial_{q^{k}}\Pi_{0}\partial_{p_{k}}\Pi_{1}$, with
Einstein's summation rule,
equals
\begin{eqnarray*}
-\partial_{q^{k}}\Pi_{0}\partial_{p_{k}}\Pi_{1}
&=&-\frac{i\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{(E_{+}-E_{-})}
(\partial_{q^{k}}\Pi_{0})^{OD}\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})^{OD}
\\
&=&\frac{i\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{(E_{+}-E_{-})}
\sigma_{3}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,,
\end{eqnarray*}
while the second term
$+(\partial_{p_{k}}\Pi_{1})(\partial_{q^{k}}\Pi_{0})$ gives the same result.\\
The third term $\Pi_{1}^{2}$ is given by
\begin{eqnarray*}
\Pi_{1}^{2}&=&-\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})^{2}}
\sigma_{3}(\partial_{q^{k}}\Pi_{0})^{OD}\sigma_{3}
(\partial_{q^{\ell}}\Pi_{0})^{OD}
\\
&=&
\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})^{2}}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
\end{eqnarray*}
Hence the diagonal second order correction is given by
\begin{eqnarray*}
&&(E_{+}-E_{-})^{2}\Pi_{2}^{D}=-(E_{+}-E_{-})^{2}\sigma_{3}G_{2}
\\
&&=
-
\left[(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})
+
(E_{+}-E_{-})(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})\sigma_{3}\right]
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})
\end{eqnarray*}
and satisfies
$$
\varepsilon^{-2\delta}\Pi_{2}^{D}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle \sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,.
$$
Consider now $\Pi_{2}^{OD}$: By referring to \eqref{eq.FOD}, $F_{2}$ can be chosen as the
off-diagonal part of
$\varepsilon^{-2}\left[H\sharp^{\varepsilon}\Pi^{(1)}-\Pi^{(1)}\sharp^{\varepsilon}H\right]$
which, according to the previous steps and \eqref{eq.Wexpanexp},
equals
\begin{multline}
\label{eq.F2}
-\frac{1}{8}\left[\partial^{2}_{p_{k},p_{\ell}}H\partial^{2}_{q^{k},q^{\ell}}\Pi_{0}
-
\partial^{2}_{q^{k},q^{\ell}}
\Pi_{0}\partial^{2}_{p_{k},p_{\ell}}H\right]
\\
+
\frac{1}{2i}\left[\partial_{p_{k}}H\partial_{q^{k}}\Pi_{1}-\partial_{q^{k}}H\partial_{p_{k}}\Pi_{1}
- \partial_{p_{k}}\Pi_{1}\partial_{q^{k}}H
+\partial_{q^{k}}\Pi_{1}\partial_{p_{k}}H\right]
\mod\mathcal{O}_{S}(\varepsilon)\,.
\end{multline}
Since
$\partial^{2}_{p_{k},p_{\ell}}H=(\partial^{2}_{p_{k},p_{\ell}}f_{\varepsilon})$
as a scalar symbol
commutes with $\partial^{2}_{q^{k}q^{\ell}}\Pi_{0}$ , the first term of \eqref{eq.F2}
vanishes.\\
Similarly the factor $\partial_{q^{k}}\Pi_{1}$ appearing in the second
term of \eqref{eq.F2}
contains three terms
\begin{multline*}
\partial_{q^{k}}\Pi_{1}=
-i\frac{\partial_{q^{k}}(E_{+}-E_{-})}{(E_{+}-E_{-})^{2}}
\sigma_{3}(\partial_{p_{\ell}}f_{\varepsilon})(\partial_{q^{\ell}}\Pi_{0})+
\frac{i(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
(\partial_{q^{k}}\sigma_{3})(\partial_{q^{\ell}}\Pi_{0})\\
+
\frac{i(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
\sigma_{3}(\partial^{2}_{q^{k}q^{\ell}}\Pi_{0})\,,
\end{multline*}
where the second one is diagonal (Remember $\sigma_{3}(q,\tau,\varepsilon)=2\Pi_{0}(q,\tau,\varepsilon)-\Id_{\field{C}} \newcommand{\nz}{\field{N}^{2}}$)\,.
Since $\partial_{p_{k}}H= \partial_{p_{k}}f_{\varepsilon}$ is
diagonal, we get
\begin{multline*}
\frac{1}{2i}\left[\partial_{p_{k}}H\partial_{q^{k}}\Pi_{1}
+\partial_{q^{k}}\Pi_{1}\partial_{p_{k}}H\right]^{OD}
\\
=
(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})\sigma_{3}
\left(\partial_{q^{k}}\left[(E_{+}-E_{-})^{-1}\partial_{q^{\ell}}\Pi_{0}\right]\right)^{OD}
\,.
\end{multline*}
In the quantity
$-\partial_{q^{k}}H \partial_{p_{k}}\Pi_{1}-\partial_{p_{k}}\Pi_{1}\partial_{q^{k}}H$,
the derivatives
$$
\partial_{p_{k}}\Pi_{1}=\frac{i(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})^{OD}
$$
are off-diagonal factors, while
$$
\partial_{q^{k}}H= (\partial_{q^{k}}E_{+})
\Pi_{0}+(\partial_{q^{k}}E_{-})(1-\Pi_{0})
+(E_{+}-E_{-})(\partial_{q^{k}}\Pi_{0})^{OD}\,.
$$
With the two equalities,
\begin{eqnarray*}
&&(\partial_{q^{k}}E_{+})\Pi_{0}\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})
+ \sigma_{3}(\partial_{q^{\ell}}\Pi_{0}) (\partial_{q^{k}}E_{+})\Pi_{0}
=(\partial_{q^{k}}E_{+})\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})\,,
\\
&&
\hspace{-0.7cm}
(\partial_{q^{k}}E_{-})(1-\Pi_{0})\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})
+ \sigma_{3}(\partial_{q^{\ell}}\Pi_{0}) (\partial_{q^{k}}E_{-})(1-\Pi_{0})
= (\partial_{q^{k}}E_{-})\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})\,,
\end{eqnarray*}
we get
$$
-\frac{1}{2i}\left[\partial_{q^{k}}H \partial_{p_{k}}\Pi_{1}+\partial_{p_{k}}\Pi_{1}\partial_{q^{k}}H\right]^{OD}
=
-\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2(E_{+}-E_{-})}(\partial_{q^{k}}(E_{+}+E_{-}))
\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})\,.
$$
This leads to
\begin{multline*}
F_{2}=
(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})
\sigma_{3}\left(\partial_{q^{k}}\left[(E_{+}-E_{-})^{-1}\partial_{q^{\ell}}\Pi_{0}\right]\right)^{OD}
\\
-\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{(E_{+}-E_{-})}(\partial_{q^{k}}(E_{+}+E_{-}))
\sigma_{3}(\partial_{q^{\ell}}\Pi_{0})\,.
\end{multline*}
and
\begin{multline*}
\Pi_{2}^{OD}=
-\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
\left(\partial_{q^{k}}\left[(E_{+}-E_{-})^{-1}\partial_{q^{\ell}}\Pi_{0}\right]\right)^{OD}
\\
+\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2(E_{+}-E_{-})^{2}}(\partial_{q^{k}}(E_{+}+E_{-}))
(\partial_{q^{\ell}}\Pi_{0})\,,
\end{multline*}
with
$$
\varepsilon^{-2\delta}\Pi_{2}^{OD}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle \sqrt{\tau'\tau''}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,.
$$
We have almost proved the
\begin{proposition}
\label{pr.proj2nd}
The pseudodifferential operator
$\hat\Pi(\varepsilon)=\hat\Pi(q,\varepsilon D_{q},\varepsilon)$ given by
\begin{eqnarray*}
&& \Pi(q,p,\varepsilon)=\Pi_{0}(q,\varepsilon)+\varepsilon
\Pi_{1}(q,p,\varepsilon)+ \varepsilon^{2}\Pi_{2}(q,p,\varepsilon)\\
\text{with}&&
\Pi_{1}(q,p,\varepsilon)=
\Pi_{1}(q,p,\varepsilon)^{OD}=\frac{i
\partial_{p_{k}}f_{\varepsilon}}{(E_{+}-E_{-})}
\sigma_{3}(\partial_{q^{k}}\Pi_{0})
\,,\\
&&\Pi_{2}(q,p,\varepsilon)=
\Pi_{2}(q,p,\varepsilon)^{D}+\Pi_{2}(q,p,\varepsilon)^{OD}\,,\\
&&
\Pi_{2}(q,p,\varepsilon)^{D}
=
-\frac{1}{(E_{+}-E_{-})^{2}}\left[(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})
\right.\\
&&\hspace{4.5cm}\left.+(E_{+}-E_{})(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})\right]
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,,\\
&&
\Pi_{2}(q,p,\varepsilon)^{OD}
= -\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
\left(\partial_{q^{k}}\left[(E_{+}-E_{-})^{-1}\partial_{q^{\ell}}\Pi_{0}\right]\right)^{OD}
\\
&&\hspace{4.5cm}
+2\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{(E_{+}-E_{-})^{3}}(\partial_{q^{k}}(E_{+}+E_{-}))
(\partial_{q^{\ell}}\Pi_{0})\,,\\
\text{and}&&
\varepsilon^{-2\delta}\Pi_{1}, \varepsilon^{-2\delta}\Pi_{2}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle \sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,,
\end{eqnarray*}
satisfies
$$
\hat \Pi\circ \hat \Pi= \hat \Pi+
\mathcal{O}(\varepsilon^{3+2\delta})
\quad,\quad
\hat \Pi^{*}=\hat \Pi \quad,\quad
\left[\hat H, \hat \Pi\right]= \mathcal{O}(\varepsilon^{3+2\delta})\,,
$$
in $\mathcal{L}(L^{2}(\field{R}^{d};\field{C}} \newcommand{\nz}{\field{N}^{2}))$\,. Moreover the estimates in
$\mathcal{L}(L^{2}(\field{R}^{d};\field{C}} \newcommand{\nz}{\field{N}^{2}))$ of
the remainder terms
do not depend on the parameter $\tau=(\tau',\tau'')\in (0,1]^{2}$ and
$\delta>0$, as soon as $\varepsilon\in (0,\varepsilon_{0})$\,.
\end{proposition}
\noindent{\bf Proof: \ }{}
The above construction gives immediately
$$
\hat \Pi\circ \hat \Pi= \hat \Pi+
\mathcal{O}(\varepsilon^{3})
\quad,\quad
\hat \Pi^{*}=\hat \Pi \quad,\quad
\left[\hat H, \hat \Pi\right]= \mathcal{O}(\varepsilon^{3})\,,
$$
The first improved estimates come from the fact that
$$
\hat \Pi\circ \hat \Pi- \hat \Pi
$$
contains only terms which are Moyal products with a $\Pi_{1}$ or a
$\Pi_{2}$ factor, with cancellations up to the $\varepsilon^{2}$
coefficient.
Both of them have seminorms of
order $\varepsilon^{2\delta}$\,.\\
For the last one, this is a similar argument after decomposing
$$
\left[\hat H, \hat \Pi\right]
=
\left[\varepsilon^{2\delta}f_{1}(\sqrt{\tau'\tau''}\varepsilon D_{q}), \hat \Pi_{0}\right]
+ \left[\hat H, \varepsilon \hat \Pi_{1}+\varepsilon^{2}\hat\Pi_{2}\right]\,.
$$
{\hfill $\qed${}
The above result can be improved after considering what
happens at step $n=3$ when
$f_{\varepsilon,\tau}(p)=\varepsilon^{2\delta}f_{1}(\sqrt{\tau'\tau''}p)$
is at most quadratic w.r.t $p$ in
some region.
Before this, let us examine the remainders of order $3$ in
$\varepsilon$.\\
\noindent$\mathbf n=3:$ The remainder term
\begin{eqnarray*}
\varepsilon^{3}G_{3}&=&\Pi^{(2)}\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}
\\
&=&
\varepsilon\left(\Pi_{0}\sharp^{\varepsilon}
\Pi_{1}+\Pi_{1}\sharp^{\varepsilon}\Pi_{0}-\Pi_{1}\right)
+\varepsilon^{2}\left(\Pi_{1}\sharp^{\varepsilon}
\Pi_{1}+\Pi_{0}\sharp^{\varepsilon}
\Pi_{2}+\Pi_{2}\sharp^{\varepsilon}\Pi_{0}-\Pi_{2}\right)\\
&&+\varepsilon^{3}(\Pi_{1}\sharp^{\varepsilon}\Pi_{2}+\Pi_{2}\sharp^{\varepsilon}\Pi_{1})
+\varepsilon^{4}\Pi_{2}\sharp^{\varepsilon}\Pi_{2}\,.
\end{eqnarray*}
Using the construction of $\Pi_{1}$ and $\Pi_{2}$ and the fact that
$\varepsilon^{-2\delta}\Pi_{1}$ and $\varepsilon^{-2\delta}\Pi_{2}$
belong to $S_{u}\Big(\frac{1}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$, the
expansion of the Moyal product \eqref{eq.Wexpan3}-\eqref{eq.Wexpanexp} tells us
\begin{equation}
\label{eq.G3}
\Pi^{(2)}\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}= \varepsilon^{3}\left[A_{G}(\partial_{q}^{2}\Pi_{0},\partial_{p}^{2}\Pi_{1},\varepsilon)
+ B_{G}(\partial_{q}\Pi_{0}, \partial_{p}\Pi_{2},\varepsilon) \right] +\varepsilon^{3+4\delta}R_{G}
\end{equation}
where $R_{G}\in S_{u}\Big(\frac{1}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$ and
$A_{G}(.,\varepsilon)$, $B_{G}(.,\varepsilon)$ have an asymptotic expansion in
terms of $\varepsilon$, of which all the terms are bilinear differential expressions
of their arguments.
For the commutator with $\hat{H}$, write
\begin{eqnarray*}
\left[H\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}\sharp^{\varepsilon}H\right]
&=&\left[f_{\varepsilon}(p)\sharp^{\varepsilon}\Pi_{0}(q)-\Pi_{0}(q)\sharp^{\varepsilon}f_{\varepsilon}(p)\right]
\\
&&+
\left[f_{\varepsilon}(p)\sharp^{\varepsilon}(\varepsilon\Pi_{1}+\varepsilon^{2}\Pi_{2})-(\varepsilon\Pi_{1}+\varepsilon^{2}\Pi_{2})\sharp^{\varepsilon}f_{\varepsilon}(p)\right]
\\
&&
\hspace{-2cm}+
\left[(E_{+}\Pi_{0}+E_{-}(1-\Pi_{0}))\sharp^{\varepsilon}\Pi^{(2)}
-\Pi^{(2)}\sharp^{\varepsilon}(E_{+}\Pi_{0}+E_{-}(1-\Pi_{0}))\right]
\end{eqnarray*}
After eliminating all the terms which are cancelled while constructing
$\Pi_{1}$ and $\Pi_{2}$, the contributions of all three terms of the
right-hand side can be analyzed. The contribution of the third term is
similar to what we got for $G_{3}$:
$$
\varepsilon^{3}\left[A_{H}(\partial_{q}^{2}(E_{\pm}\Pi_{0}), \partial_{p}^{2}\Pi_{1},\varepsilon)
+
B_{H}(\partial_{q}(E_{\pm}\Pi_{0}), \partial_{p}\Pi_{2},\varepsilon)\right]
+\varepsilon^{3+4\delta}R_{H,1}
$$
with $R_{H,1}\in S_{u}\Big(\frac{1}{\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$\,.
With the uniform estimate of $\varepsilon^{-2\delta}\Pi_{1}$,
$\varepsilon^{-2\delta}\Pi_{2}$ and
$\varepsilon^{-2\delta}f_{\varepsilon}$, the contribution of the
second term is estimated as $\varepsilon^{3+4\delta}R_{H,2}$ with
$R_{H,2}\in S_{u}\Big(\frac{1}{\langle\sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$\,. The contribution of the first term is
$\varepsilon^{3}C_{H}(\partial_{p}^{3}f_{\varepsilon}, \partial_{q}^{3}\Pi_{0},\varepsilon)$
and we get
\begin{multline}
\label{eq.almF3}
\left[H\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}\sharp^{\varepsilon}H\right]
=
\varepsilon^{3}\left[
C_{H}(\partial_{p}^{3}f_{\varepsilon}, \partial_{q}^{3}\Pi_{0},\varepsilon)
+
\right.
\\
\left.A_{H}(\partial_{q}^{2}(E_{\pm}\Pi_{0}), \partial_{p}^{2}\Pi_{1},\varepsilon)
+
B_{H}(\partial_{q}(E_{\pm}\Pi_{0}), \partial_{p}\Pi_{2},\varepsilon)
\right]
+\varepsilon^{3+4\delta}R_{H}
\end{multline}
where the expansions of $A_{H},B_{H}$ and $C_{H}$ w.r.t to $\varepsilon$ have
terms which are bilinear differential expression of their arguments,
and the remainder $R_{H}$ belongs to
$S_{u}\Big(\frac{1}{\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$\,.
\begin{proposition}
\label{pr.preci}
With
$f_{\varepsilon}(p)=\varepsilon^{2\delta}f_{1}(\sqrt{\tau'\tau''}p)$,
assume that the third differential $\partial_{p}^{3}f_{1}$ vanishes
in $\left\{|p|\leq r\right\}$ and fix $r'\in (0,r)$\,. Then the remainders of
Proposition~\ref{pr.proj2nd} equal
\begin{eqnarray}
\label{eq.Piprec}
&& \hat{\Pi}\circ \hat{\Pi}-\hat{\Pi}=
\varepsilon^{3+2\delta}\hat{R}_{G,1}+
\varepsilon^{3+4\delta}\hat{R}_{G,2}\\
\label{eq.Hprec}
&& \left[\hat{H},\hat{\Pi}\right]= \varepsilon^{3+2\delta}\hat{R}_{H,1}+
\varepsilon^{3+4\delta}\hat{R}_{H,2}
\end{eqnarray}
where $R_{G,1\text{or}2}$ belong to $OpS_{u}\Big(\frac{1}{\langle\sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$, $R_{H,1\text{or}2}$ belong to $ OpS_{u}\Big(\frac{1}{\sqrt{\tau'\tau''}p\rangle^{\infty}}, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$ and
$R_{G\text{or}H,1}\equiv 0$ in
$\left\{|\sqrt{\tau'\tau''}p|<r'\right\}$\,. Those estimates are
uniform for $\tau\in (0,1]^{2}$, $\delta\in [0,\delta_{0}]$ and
$\varepsilon\in (0,\varepsilon_{0})$\,.
\end{proposition}
\noindent{\bf Proof: \ }{}
After noticing that the symbol $\Pi_{1}$ is a linear expression in
$\partial_{p}f_{\varepsilon}$ while the symbol
$\Pi_{2}$ is the sum of a linear expression of $\partial_{p}^{2}f_{\varepsilon}$
and quadratic expression in $\partial_{p}f_{\varepsilon}$, the
identities \eqref{eq.G3} and \eqref{eq.almF3} imply
\begin{eqnarray*}
\Pi^{(2)}\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}
&=& \varepsilon^{3+2\delta+N}R_{N}+ \varepsilon^{3+4\delta}R_{G}\,,\\
\left[H\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}\sharp^{\varepsilon}H\right]
&=&
\varepsilon^{3+2\delta+ N}R_{N}+ \varepsilon^{3+4\delta}R_{H}\,,
\end{eqnarray*}
for an arbitrary large $N\in\nz$,
in $\left\{\sqrt{\tau'\tau''}|p|< r\right\}$\,. Choose\footnote{Here
the estimates become $\delta$-dependent, because a large $\delta$
requires a large $N$. It is uniformly controlled when $\delta\leq \delta_{0}$.} $N\geq
2\delta$ and take a cut-off function $\chi\in
\mathcal{C}^{\infty}_{0}\left(\left\{|p|<r\right\}\right)$ such that
$\chi\equiv 1$ in a neighborhood $\left\{|p|\leq r'\right\}$\,. Writing
for the symbol
$$
S=
\Pi^{(2)}\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}\quad\text{or}\quad
S=\left[H\sharp^{\varepsilon}\Pi^{(2)}-\Pi^{(2)}\sharp^{\varepsilon}H\right]\,,$$
$$
S=S\times\chi(\sqrt{\tau'\tau''}p)+ S \times(1-\chi(\sqrt{\tau'\tau''}p))\,,
$$
yields the result.
{\hfill $\qed${}
\subsection{Unitaries and effective Hamiltonian}
\label{se.projunit}
We strengthen a little bit the assumptions
\eqref{eq.hypEPi1}-\eqref{eq.hypEPi2}, with the condition
\begin{eqnarray}
\label{eq.hypu0}
&&\Pi_{0}(q,\tau,\varepsilon)=u_{0}(q,\tau,\varepsilon)P_{+}u_{0}(q,\tau,\varepsilon)^{*}
\\
\nonumber
\text{with}
&&
P_{+}=
\begin{pmatrix}
1&0\\0&0
\end{pmatrix}
\,,\; u_{0}=(u_{0}^{*})^{-1}\in S(1,g_{q,\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\end{eqnarray}
fulfilled in our example. The operator $\hat{u}_{0}$ is nothing but
the local unitary transformation $u_{0}(q,\tau)$, on $L^{2}(\field{R}^{d};\field{C}} \newcommand{\nz}{\field{N}^{2})$\,.\\
With the approximate projection $\hat\Pi=\hat{\Pi}(q,\varepsilon
D_{q},\tau,\varepsilon)$ given in Proposition~\ref{pr.proj2nd},
Proposition~\ref{pr.appproj}
tells us that a true
orthogonal projection $\hat{P}$ can be associated when
$\varepsilon_{0}$ is chosen small enough, by taking
\begin{equation}
\label{eq.defPeps}
\hat{P}=\frac{1}{2i\pi}\int_{|z-1|=1/2}(z-\hat\Pi)^{-1}~dz\,,
\end{equation}
with
\begin{eqnarray}
\label{eq.hatP}
&&
\hat{P}=P(q,\varepsilon D_{q},\tau,\varepsilon)\quad,\quad
P\in S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,\\
\label{eq.PPi}
&&
\hspace{-3cm}
P(q,p,\tau,\varepsilon)-\Pi(q,p,\tau,\varepsilon)=\varepsilon^{3+2\delta}R_{1}(q,p,\tau,\varepsilon)
+\varepsilon^{3+4\delta}R_{2}(q,p,\tau,\varepsilon)\,,\;\\
\nonumber
\text{with}&&
R_{1},R_{2}\in S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle \sqrt{\tau'\tau''}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,,\\
&&\nonumber
\hat{P}\circ
\hat{P}=\hat{P}=\hat{P}^{*}\,,
\\
&&
\label{eq.comHP}
\left[\hat{H},\hat{P}\right]=\varepsilon^{3+2\delta}
\hat{C}_{1}(\tau,\varepsilon)+\varepsilon^{3+4\delta}\hat{C}_{2}(\tau,\varepsilon)\,,\\
\nonumber\quad\text{with}&&
C_{1},C_{2}\in S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,\\
\text{and}
&&
\label{eq.Pu0P+}
\left\|\hat{P}-\hat{u}_{0}P_{+}\hat{u}_{0}^{*}\right\|_{\mathcal{L}(L^{2})}\leq C\varepsilon\,.
\end{eqnarray}
For a general $f_{1}\in \mathcal{C}^{\infty}_{0}$, $R_{2}$ and $C_{2}$ are included in the main
remainder term. When $f_{1}$ is quadratic in $\left\{|p|<r\right\}$,
then one can assume that $R_{1}$ and $C_{1}$ vanishes in
$\left\{\sqrt{\tau'\tau''}|p|<r'\right\}$ for $r'<r$, according to
Proposition~\ref{pr.preci} and Proposition~\ref{pr.appproj}\\
Instead of constructing unitaries between $\hat{P}(\tau,\varepsilon)$ and $P_{+}$
by the induction presented in
\cite{PST} and similar to \eqref{eq.defGn1},
\eqref{eq.Pin1D},\eqref{eq.defFn1},\eqref{eq.Pin1OD}, we use like in \cite{MaSo} Nagy's formula
(\cite{NeSo}, \cite{MaSo}, \cite{PST})
\begin{eqnarray}
\nonumber
&&P_{2}=WP_{1}W^{*}\quad,\quad W^{*}W=WW^{*}=1\,,\\
\text{with}
&&
\label{eq.nagy}
W=(1-(P_{2}-P_{1})^{2})^{-1/2}\left[P_{2}P_{1}+(1-P_{2})(1-P_{1})\right]\,,\\
\nonumber
\text{when}&&
P_{j}=P_{j}^{2}=P_{j}^{*}\quad \text{for}\, j=1,2\,,\quad
\text{and}\,\|P_{2}-P_{1}\|_{\mathcal{L}}< 1\,,
\end{eqnarray}
easier to handle for direct second order computations in our case.
\begin{proposition}
\label{pr.projunit}
With the definitions \eqref{eq.hypu0} and \eqref{eq.defPeps} after
Proposition~\ref{pr.proj2nd}, there exists a unitary operator
$\hat{U}$ on $L^{2}(\field{R};\field{C}} \newcommand{\nz}{\field{N}^{2})$ such that
\begin{eqnarray*}
&& \hat P=\hat{U}P_{+}\hat{U}^{*}\,,\\
&& \hat{U}=U(q,\varepsilon D_{q},\tau,\varepsilon)\quad,\quad
U\in S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\,,
\\
\text{with}&&U(q,p,\tau,\varepsilon)=u_{0}(q,\tau,\varepsilon)+\varepsilon u_{1}(q,p,\tau,\varepsilon)+
\varepsilon^{2}u_{2}(q,p,\tau,\varepsilon)\,,
\\
&&
u_{1}(q,p,\varepsilon)=-\frac{i\partial_{p_{k}}f_{\varepsilon}}{(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})
u_{0}\,,\\
\text{and}&& \varepsilon^{-2\delta}u_{1},
\varepsilon^{-2\delta}u_{2}\in S_{u}\Big(
\frac{1}{\langle\sqrt{\frac{\tau'}{\tau''}}\rangle\langle\sqrt{\tau'\tau''}p\rangle^{\infty}},
g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)\,.
\end{eqnarray*}
Moreover, when $f_{1}$ is a quadratic function in
$\left\{|p|<r\right\}$ and $r'$ is fixed in $ (0,r)$, the term $u_{2}$ can be decomposed into
$$
u_{2}(q,p,\tau,\varepsilon)=
\varepsilon^{2\delta}v_{2}(q,p,\tau,\varepsilon)+
\varepsilon^{4\delta}\widetilde{v_{2}}(q,p,\tau,\varepsilon)
$$
where $v_{2}$ does not depend on $p$ in
$\left\{\sqrt{\tau'\tau''}|p|<r'\right\}$ and $v_{2},
\widetilde{v}_{2}$ belong to the symbol class
$S_{u}\Big(\frac{1}{\langle\sqrt{\frac{\tau'}{\tau''}}\rangle\langle\sqrt{\tau'\tau''}p\rangle^{\infty}},
g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$\,.
\end{proposition}
\noindent{\bf Proof: \ }{}
The notation $\hat{R}$ will denote a generic remainder term of
the form $\hat{R}=R(q,\varepsilon D_{q},\tau,\varepsilon)$ with
$R\in S_{u}\Big(\frac{1}{\langle
\sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle\sqrt{\tau'\tau''}
p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$\,. The notation
$\hat{\underline{R}}$ is used for a symbol $\underline{R}$, like $R$ but which vanishes
around $\left\{\sqrt{\tau'\tau''}|p|<r'\right\}$\,.
We apply Nagy's formula \eqref{eq.nagy} with
$P_{1}=\hat{u}_{0}P_{+}\hat{u}_{0}^{*}=\Pi_{0}(q,\tau,\varepsilon)$
and $P_{2}=\hat{P}(\tau,\varepsilon)$ with
$$
\hat P=\Pi_{0}(q,\tau,\varepsilon)+\varepsilon \Pi_{1}(q,\varepsilon
D_{q},\tau,\varepsilon)+ \varepsilon^{2}\Pi_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)+ \varepsilon^{3+2\delta}\underline{\hat{R}}+\varepsilon^{3+4\delta}\hat{R}\,.
$$
In the expression of $\hat{W}(\varepsilon)$ given by \eqref{eq.nagy},
the first factor is nothing but
$$
(1-(\hat{P}-\Pi_{0}(q,\tau,\varepsilon))^{2})^{-1/2}=1+
\frac{\varepsilon^{2}}{2}(\Pi_{1})^{2}(q,\varepsilon
D_{q},\tau,\varepsilon)
+
\varepsilon^{3+4\delta}\hat{R}\,,
$$
owing to $P-\Pi_{0}=\mathcal{O}_{S}(\varepsilon^{2\delta})$\,.
In the factor $[P_{2}P_{1}+(1-P_{2})(1-P_{1})]$, the first term equals
\begin{multline*}
\hat{P}\circ \Pi_{0}(q,\varepsilon)=
\Pi_{0}(q,\varepsilon)+\varepsilon\Pi_{1}(q,\tau,\varepsilon
D_{q},\tau,\varepsilon)\circ \Pi_{0}(q,\tau,\varepsilon)
\\
+
\varepsilon^{2}\Pi_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)\circ \Pi_{0}(q,\tau,\varepsilon)
+\varepsilon^{3+2\delta}\underline{\hat{R}}+\varepsilon^{3+4\delta}\hat{R}\,,
\end{multline*}
while the second term is
\begin{multline*}
(1-\hat{P})\circ(1-\Pi_{0}(q,\tau,\varepsilon))=(1-\Pi_{0}(q,\tau,\varepsilon))
-\varepsilon\Pi_{1}(q,\varepsilon
D_{q},\tau,\varepsilon)\circ (1-\Pi_{0}(q,\tau,\varepsilon))
\\
-
\varepsilon^{2}\Pi_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)\circ (1-\Pi_{0}(q,\tau\varepsilon))
+\varepsilon^{3+2\delta}\underline{\hat{R}}+\varepsilon^{3+4\delta}\hat{R}\,.
\end{multline*}
Hence we get
\begin{multline*}
\hat{W}=
[P_{2}P_{1}+(1-P_{2})(1-P_{1})]=1+\varepsilon\Pi_{1}(q,\varepsilon
D_{q},\tau,\varepsilon)\circ
\sigma_{3}(q,\tau)
\\
+\varepsilon^{2}\Pi_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)\circ \sigma_{3}(q,\tau)
+\varepsilon^{3+2\delta}\underline{\hat{R}}+\varepsilon^{3+4\delta}\hat{R}\,.
\end{multline*}
The operator $\hat{U}(\varepsilon)$ is given by
$\hat{U}(\varepsilon)=\hat{W}(\varepsilon)\circ\hat{u}_{0}$\,.
The
semiclassical calculus recalled in
\eqref{eq.Wexpan3}-\eqref{eq.Wexpanexp} yields the result.
In the decomposition of $u_{2}$, when $f_{1}$ is quadratic around $0$,
the terms which are linear in
$\varepsilon^{2\delta}$ come with the second derivative of
$f_{\varepsilon}$, which does not depend on $p$\,.
{\hfill $\qed${}
\begin{proposition}
\label{pr.hameff}
Introduce the
notation for $k\in \left\{1,\ldots,d\right\}$
\begin{equation*}
u_{0}^{*}(\partial_{q^{k}}u_{0})=-i
\begin{pmatrix}
A_{k}&X_{k}\\
\overline{X_{k}} & -A_{k}
\end{pmatrix}\,.
\end{equation*}
If $\hat{U}$ is the unitary operator introduced in
Proposition~\ref{pr.projunit}, the conjugated Hamiltonian
$\hat{U}(\varepsilon)^{*}\hat{H}(\varepsilon)\hat{U}(\varepsilon)$ equals
\begin{equation}
\label{eq.almdiag}
\hat{U}^{*}\hat{H}\hat{U}
=
\begin{pmatrix}
\hat{h}_{+}&0\\
0& \hat{h}_{-}
\end{pmatrix}
+ \varepsilon^{3+2\delta}\hat{R}_{1}+\varepsilon^{3+4\delta}\hat{R}_{2}
\end{equation}
where the remainder terms $\hat{R}_{1,2}=R_{1,2}(q,\varepsilon
D_{q},\tau,\varepsilon)$ belong to
$OpS_{u}\left(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\right)$ and additionally
$R_{1}\equiv 0$ in $\left\{\sqrt{\tau'\tau''}|p|<r'\right\}$ when $f_{1}$
is quadratic in $\left\{|p|<r\right\}$, with $r'<r$\,.\\
The symbol $h_{+}$ and $h_{-}$ are given by
\begin{eqnarray}
\nonumber
h_{+}&=&
f_{\varepsilon}(p)+E_{+}(q)-\varepsilon
(\partial_{p}f_{\varepsilon}).A
+
\frac{\varepsilon^{2}}{2}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})A_{k}A_{\ell}
\\
\label{eq.h+}
&&\hspace{1cm}
+
\frac{\varepsilon^{2}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}
{2}X_{k}\overline{X}_{\ell}
+
\frac{\varepsilon^{2}(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}
{E_{+}-E_{-}}X_{k}\overline{X}_{\ell}
\,,
\\
\nonumber
h_{-}&=&
f_{\varepsilon}(p)+E_{-}(q)+\varepsilon
(\partial_{p}f_{\varepsilon}).A
+
\frac{\varepsilon^{2}}{2}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})A_{k}A_{\ell}
\\
\label{eq.h-}
&&\hspace{1cm}
+\frac{\varepsilon^{2}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{2}
\overline{X_{k}}X_{\ell}
-
\frac{\varepsilon^{2}(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}
{E_{+}-E_{-}}\overline{X_{k}}X_{\ell}\,.
\end{eqnarray}
\end{proposition}
\noindent{\bf Proof: \ }{}
From the semiclassical calculus, we already know that
$\hat{U}^{*}\hat{H}\hat{U}$ is a
semiclassical operator with a symbol in
$S_{u}(\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle,
g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,. Its off-diagonal part equals
\begin{eqnarray*}
(1-P_{+})\hat{U}^{*}\hat{H}\hat{U}P_{+}
&+&
P_{+}\hat{U}^{*}\hat{H}\hat{U}(1-P_{+})
\\
&& =
\hat{U}\left[
\hat{P},\left[\hat{P},\hat{H}\right]\right] \hat{U}\,.
\end{eqnarray*}
The almost diagonal form \eqref{eq.almdiag}
of $\hat{U}^{*}\hat{H}\hat{U}$
is then a consequence of \eqref{eq.comHP}\,.\\
For the second result, it is necessary to compute the diagonal part of the symbol
$U^{*}\sharp^{\varepsilon}H\sharp^{\varepsilon}U$
up to $\mathcal{O}(\varepsilon^{3+2\delta})$ in
$S_{u}(1,g_{\tau};\mathcal{M}_{2})$\,.
Let us compute the diagonal part of
$$
\mathcal{B}:=(u_{0}^{*}+\varepsilon u_{1}^{*}+\varepsilon^{2}u_{2}^{*})\sharp^{\varepsilon}H\sharp^{\varepsilon}(u_{0}+\varepsilon
u_{1}+\varepsilon^{2}u_{2})\,,
$$
or equivalently $(u_{0}\mathcal{B}u_{0}^{*})^{D}$ with
our notations.\\
Since $u_{0}=u_{0}(q,\tau,\varepsilon)$ and
$H=\varepsilon^{2\delta}f_{1}(\sqrt{\tau'\tau''}p)+\mathcal{V}(q,\tau,\varepsilon)$
\,, the first Moyal
product equals according to \eqref{eq.Wexpan3}-\eqref{eq.Wexpanexp},
\begin{multline*}
(u_{0}^{*}+\varepsilon u_{1}^{*}+\varepsilon^{2}u_{2}^{*})\sharp^{\varepsilon}H
=
u_{0}^{*}H+\varepsilon u_{1}^{*}H+\varepsilon^{2}u_{2}^{*}H
\\
+\frac{\varepsilon}{2i}\left\{u_{0}^{*},H\right\}+
\frac{\varepsilon^{2}}{2i}\left\{u_{1}^{*},H\right\}
-\frac{\varepsilon^{2}}{8}(\partial^{2}_{q^{k}q^{\ell}}u_{0}^{*})(\partial^{2}_{p_{k}p_{\ell}}H)
+ \varepsilon^{3+2\delta}R
\\
=u_{0}^{*}H+\varepsilon
\left[u_{1}^{*}H-
\frac{\partial_{p_{k}}f_{\varepsilon}}{2i}\partial_{q^{k}}u_{0}^{*}\right]
+\varepsilon^{2}
\left[u_{2}^{*}H
+ \frac{1}{2i}\left\{u_{1}^{*},H\right\}
\right.
\\
\left.
-\frac{1}{8}(\partial^{2}_{q^{k}q^{\ell}}u_{0}^{*})(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})
\right]+ \varepsilon^{3+2\delta}\underline{R}+\varepsilon^{3+4\delta}R\,,
\end{multline*}
where $R$ and $\underline{R}$ denote generic element of
$S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, with the additional property
that $\underline{R}$ vanishes in
$\left\{\sqrt{\tau'\tau''}|p|<r'\right\}$ when $f_{1}$ is quadratic
in $\left\{|p|\leq r\right\}$ with $r'<r$. The reason for the possible
decomposition of the remainder, comes again from the fact that the
third order remainder term, proportional to $\varepsilon^{3+2\delta}$,
arises with the third derivative of $f_{\varepsilon}$, the second
derivative w.r.t $p$ of $u_{1}$ and the first derivative w.r.t $p$ of
$u_{2}$\,.\\
In the same way, the complete expression of
$\mathcal{B}$ is given by
\begin{eqnarray*}
&&\hspace{-0.5cm}\mathcal{B}(\varepsilon)=u_{0}^{*}Hu_{0}+\varepsilon\left[u_{0}^{*}H
u_{1}+u_{1}^{*}Hu_{0}-\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}(\partial_{q^{k}}u_{0}^{*})u_{0} +
\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}u_{0}^{*}(\partial_{q^{k}}u_{0}) \right]
\\
&&\;
+\varepsilon^{2}\left[u_{2}^{*}H u_{0}+
u_{0}^{*}H u_{2}
+\frac{1}{2i}\left\{u_{0}^{*}H,u_{1}\right\}+
\frac{1}{2i}\left\{u_{1}^{*}H, u_{0}\right\}
+\frac{1}{4}\left\{(\partial_{p_{k}}f_{\varepsilon})(\partial_{q^{k}}u_{0}^{*}),
u_{0}\right\}
\right.\\
&&\quad
+u_{1}^{*}Hu_{1}-\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}(\partial_{q^{k}}u_{0}^{*})u_{1}
+\frac{1}{2i}\left\{u_{1}^{*},H\right\}u_{0}
\\
&&\quad
\left.
-\frac{1}{8}(\partial^{2}_{q^{k}q^{\ell}}u_{0}^{*})(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}))u_{0}
-\frac{1}{8}u_{0}^{*}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})(\partial^{2}_{q^{k}q^{\ell}}u_{0})
\right]+\varepsilon^{3+2\delta}\underline{R}+\varepsilon^{3+4\delta}R\,.
\end{eqnarray*}
By recalling that $(u_{1}u_{0}^{*})^{D}=0$ by
Proposition~\ref{pr.projunit}, we get
\begin{multline*}
(u_{0}\mathcal{B}(\varepsilon)u_{0}^{*})^{D}
=
H+\varepsilon \frac{i}{2}
(\partial_{p_{k}}f_{\varepsilon})\left[u_{0}(\partial_{q^{k}}u_{0}^{*})
-(\partial_{q^{k}}u_{0})u_{0}^{*}
\right]^{D}
\\
+\varepsilon^{2} \mathcal{B}_{2}^{D}+\varepsilon^{3+2\delta}\underline{R}+\varepsilon^{3+4\delta}R\,.
\end{multline*}
where $\mathcal{B}_{2}^{D}$ is made of several terms to be analyzed.
We need the relations
\begin{eqnarray}
\label{eq.deru0u0}
&&\partial_{q^{j}}u_{0}^{*}=-u_{0}^{*}(\partial_{q^{j}}u_{0})u_{0}^{*}\,,\\
\nonumber
&&\partial_{q^{j}q^{j'}}^{2}u_{0}^{*}
=u_{0}^{*}(\partial_{q^{j'}}u_{0})
u_{0}^{*}(\partial_{q^{j}}u_{0})u_{0}^{*}
+
u_{0}^{*}(\partial_{q^{j}}u_{0})
u_{0}^{*}(\partial_{q^{j'}}u_{0})u_{0}^{*}
-u_{0}^{*}(\partial^{2}_{q^{j}q^{j'}}u_{0})u_{0}^{*}
\\
\label{eq.der2u0u0}
&&\qquad=
-(\partial_{q^{j'}}u_{0}^{*})(\partial_{q^{j}}u_{0})u_{0}^{*}
-(\partial_{q^{j}}u_{0}^{*})(\partial_{q^{j'}}u_{0})u_{0}^{*}
-u_{0}^{*}(\partial^{2}_{q^{j}q^{j'}}u_{0})u_{0}^{*}\,,
\,\\
\label{eq.deunitOD}
&&\text{and}\qquad
\left[u_{0}(\partial_{q}u_{0}^{*})\right]^{OD}=- (\partial_{q}\Pi_{0})\sigma_{3}\,,
\end{eqnarray}
coming from $u_{0}^{*}u_{0}=1$ and the differentiation of $u_{0}^{*}\Pi_{0}u_{0}=P_{+}$\,.\\
For example, the first one simplifies the
$\mathcal{O}(\varepsilon)$-term into
$$
(u_{0}\mathcal{B}(\varepsilon)u_{0}^{*})^{D}
=
H-\varepsilon i
(\partial_{p_{k}}f_{\varepsilon})
\left[(\partial_{q^{k}}u_{0})u_{0}^{*}
\right]^{D}
+\varepsilon^{2} \mathcal{B}_{2}^{D}
+
\varepsilon^{3+2\delta}\underline{R}+\varepsilon^{3+4\delta}R\,.
$$
Many cancellations appear after assembling all the terms in
$\mathcal{B}_{2}^{D}$\,. We need accurate expressions for all of them:
\begin{itemize}
\item By using again $(u_{1}u_{0}^{*})^{D}=0$,
the term $\left[u_{0}u_{2}^{*}H+H
u_{2}u_{0}^{*}+u_{0}u_{1}^{*}Hu_{1}u_{0}^{*}\right]^{D}$
equals
\begin{multline*}
(f_{\varepsilon}+E_{+})\Pi_{0}(u_{0}u_{2}^{*}+u_{2}u_{0}^{*})\Pi_{0}+
(f_{\varepsilon}+E_{-})(1-\Pi_{0})(u_{0}u_{2}^{*}+u_{2}u_{0}^{*})(1-\Pi_{0})
\\
+
(f_{\varepsilon}+E_{-})\Pi_{0}(u_{0}u_{1}^{*}u_{1}u_{0}^{*})\Pi_{0}+
(f_{\varepsilon}+E_{+})(1-\Pi_{0})(u_{0}u_{1}^{*}u_{1}u_{0}^{*})(1-\Pi_{0})\,.
\end{multline*}
In the relation
$$
u_{0}^{*}u_{2}+u_{2}^{*}u_{0}+u_{1}^{*}u_{1}+\frac{1}{2i}\left\{u_{0}^{*},u_{1}\right\}
+
\frac{1}{2i}\left\{u_{1}^{*},u_{0}\right\}=\varepsilon^{1+2\delta}
\underline{R_{1}}
+\varepsilon^{1+4\delta}R_{1}\,,
$$
the remainder terms satisfy $R_{1},\underline{R}_{1}\in
S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau'}{\tau''}}q'\rangle\langle
\sqrt{\tau'\tau''}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$,
with the same convention as for $R,\underline{R}$\,. This identity is
obtained by writing that the $\mathcal{O}(\varepsilon^{2})$ remainder
of $U^{*}\sharp^{\varepsilon}U-1$ vanishes and
by noticing that the remainder $\underline{R_{1}}+\varepsilon^{2\delta}R_{1}$
involves second derivatives of $u_{1}$ w.r.t $p$ and first derivatives
of $u_{2}$ w.r.t $p$\,.
We obtain
\begin{multline*}
\left[u_{0}u_{2}^{*}H+H
u_{2}u_{0}^{*}+u_{0}u_{1}^{*}Hu_{1}u_{0}^{*}\right]^{D}
=(E_{-}-E_{+})(u_{0}u_{1}^{*}u_{1}u_{0}^{*})^{D}\sigma_{3}\\
-\frac{1}{2i}H\left[u_{0}\left\{u_{0}^{*},u_{1}\right\}u_{0}^{*}+u_{0}\left\{u_{1}^{*},u_{0}\right\}u_{0}^{*}\right]^{D}
+\varepsilon^{1+2\delta} \underline{\tilde{R}} + \varepsilon^{1+4\delta}\tilde{R}\,.
\end{multline*}
with $\tilde{R},\underline{\tilde{R}}\in
S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, again with the same convention.\\
Again with
$(\partial_{p}u_{1})u_{0}^{*}=(\partial_{p}u_{1}u_{0}^{*})^{OD}=\frac{(\partial^{2}_{p,p_{k}}f)}{i(E_{+}-E_{-})}
\partial_{q^{k}}\Pi_{0}$ and
\eqref{eq.deunitOD}, the last factor is
\begin{eqnarray}
\nonumber
&&\hspace{-1cm}
\left[u_{0}\left\{u_{0}^{*},u_{1}\right\}u_{0}^{*}
+u_{0}\left\{u_{1}^{*},u_{0}\right\}u_{0}^{*}\right]^{D}
=
-(u_{0}\partial_{q^{\ell}}u_{0}^{*})^{OD}
\frac{(\partial^{2}_{p_{\ell}p_{k}}f_{\varepsilon})}{i(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})
\\
\nonumber
&&\hspace{4cm}
-
\frac{(\partial^{2}_{p_{\ell}p_{k}}f_{\varepsilon})}{i(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})((\partial_{q^{\ell}}u_{0})u_{0}^{*})^{OD}
\\
\label{eq.poissdiag}
&&\hspace{3cm}=-2\frac{(\partial^{2}_{p_{\ell}p_{k}}f_{\varepsilon})}{i(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}\,.
\end{eqnarray}
With
$u_{1}u_{0}^{*}=\frac{(\partial_{p_{k}}f_{\varepsilon})}{i(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})$\,,
we have proved
\begin{multline}
\label{eq.21}
\left[u_{0}u_{2}^{*}H+H
u_{2}u_{0}^{*}+u_{0}u_{1}^{*}Hu_{1}u_{0}^{*}\right]^{D}
=-\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{(E_{+}-E_{-})}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}
\\
-H\frac{(\partial^{2}_{p_{\ell}p_{k}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}
+ \varepsilon^{1+2\delta} \underline{\tilde{R}_{1}}+\varepsilon^{1+4\delta}\tilde{R}_{1}\,.
\end{multline}
\item The term
$$
\left[\frac{1}{2i}u_{0}\left\{u_{0}^{*}f_{\varepsilon},u_{1}\right\}u_{0}^{*}+
\frac{1}{2i}u_{0}\left\{u_{1}^{*}f_{\varepsilon},u_{0}\right\}u_{0}^{*}-\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}
u_{0}(\partial_{q^{k}}u_{0}^{*})u_{1}u_{0}^{*}\right]^{D}
$$
equals
\begin{eqnarray*}
&&
\frac{1}{2i}f_{\varepsilon}\left[u_{0}\left\{u_{0}^{*},u_{1}\right\}u_{0}^{*}+u_{0}\left\{u_{1}^{*},u_{0}\right\}u_{0}^{*}\right]^{D}
\\
&&
+\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}
\left[(\partial_{q^{k}}u_{1})u_{0}^{*}+u_{0}u_{1}^{*}(\partial_{q^{k}}u_{0})u_{0}^{*}
-u_{0}(\partial_{q^{k}}u_{0}^{*})u_{1}u_{0}^{*}\right]^{D}\,.
\end{eqnarray*}
The diagonal part of $(\partial_{q^{k}}u_{1})u_{0}^{*}$ is
$$
\left[(\partial_{q^{k}}u_{1})u_{0}^{*}\right]^{D}=\frac{\partial_{p_{\ell}}f_{\varepsilon}}{i(E_{+}-E_{-})}
(\partial^{2}_{q^{k}q^{\ell}}\Pi_{0})^{D}\,.
$$
But differentiating the relation
$(\partial_{q^{k}}\Pi_{0})\Pi_{0}+\Pi_{0}(\partial_{q^{\ell}}\Pi_{0})=\partial_{q^{k}}\Pi_{0}$
w.r.t $q^{\ell}$
leads to
\begin{equation}
\label{eq.dersecdiag}
(\partial^{2}_{q^{k}q^{\ell}}\Pi_{0})^{D}=-(\partial_{q^{k}}\Pi_{0}\partial_{q^{\ell}}\Pi_{0}
+ \partial_{q^{\ell}}\Pi_{0}\partial_{q^{k}}\Pi_{0})\sigma_{3}\,.
\end{equation}
With
\eqref{eq.deunitOD} and $(u_{1}u_{0}^{*})=(u_{1}u_{0}^{*})^{OD}=\frac{\partial_{p_{\ell}}f_{\varepsilon}}{i(E_{+}-E-)}\partial_{q^{\ell}}\Pi_{0}$, we obtain
\begin{eqnarray*}
&&\hspace{-1cm}\left[u_{0}u_{1}^{*}(\partial_{q^{k}}u_{0})u_{0}^{*}
-u_{0}(\partial_{q^{k}}u_{0}^{*})u_{1}u_{0}^{*}\right]^{D}
=
[-u_{0}u_{1}^{*}\sigma_{3}(\partial_{q^{k}}\Pi_{0})+(\partial_{q^{k}}\Pi_{0})\sigma_{3}u_{1}u_{0}^{*}]
\\
&&
\hspace{2cm}=
-\frac{(\partial_{p_{\ell}}f_{\varepsilon})}{i(E_{+}-E_{-})}
\left[(\partial_{q^{\ell}}\Pi_{0})(\partial_{q^{k}}\Pi_{0})
+(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\right]\sigma_{3}\,.
\end{eqnarray*}
We have found
\begin{multline}
\label{eq.22}
\left[\frac{1}{2i}u_{0}\left\{u_{0}^{*}f_{\varepsilon},u_{1}\right\}u_{0}^{*}+
\frac{1}{2i}u_{0}\left\{u_{1}^{*}f_{\varepsilon},u_{0}\right\}u_{0}^{*}-\frac{(\partial_{p_{k}}f_{\varepsilon})}{2i}
u_{0}(\partial_{q^{k}}u_{0}^{*})u_{1}u_{0}^{*}\right]^{D}\\
=
\frac{f_{\varepsilon}(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}
+
2\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}\,.
\end{multline}
\item The diagonal part of
$\frac{1}{2i}u_{0}\left[\left\{u_{0}^{*}\mathcal{V},u_{1}\right\}
+\left\{u_{1}^{*}\mathcal{V},u_{0}\right\}\right]u_{0}^{*}$ equals
$$
\frac{1}{2i}\left[-u_{0}(\partial_{q^{k}}u_{0}^{*})\mathcal{V}(\partial_{p_{k}}u_{1})u_{0}^{*}
-(\partial_{q^{k}}\mathcal{V})(\partial_{p_{k}}u_{1})u_{0}^{*} + u_{0}(\partial_{p_{k}}u_{1}^{*})\mathcal{V}(\partial_{q^{k}}u_{0})u_{0}^{*}\right]^{D}\,.
$$
By using \eqref{eq.deunitOD} with
\begin{eqnarray*}
&&(\partial_{p_{k}}u_{1})u_{0}^{*}=\left[(\partial_{p_{k}}u_{1})u_{0}^{*}\right]^{OD}
=\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{i(E_{+}-E_{-})}(\partial_{q^{\ell}}\Pi_{0})\,,
\\
&&
(\partial_{q}\mathcal{V})^{OD}=(E_{+}-E_{-})(\partial_{q}\Pi_{0})\,,
\\
\text{and}&&
E_{-}\Pi_{0}+E_{+}(1-\Pi_{0})=\mathcal{V}+(E_{-}-E_{+})\sigma_{3}\,,
\end{eqnarray*}
it becomes
\begin{multline*}
\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})
+
\frac{1}{2i}\mathcal{V}\left[u_{0}\left\{u_{0}^{*},u_{1}\right\}u_{0}^{*}
+u_{0}\left\{u_{1}^{*},u_{0}\right\}u_{0}^{*}\right]^{D}
\\
+\frac{1}{2i}
(E_{-}-E_{+})\left[u_{0}\left\{u_{0}^{*},u_{1}\right\}u_{0}^{*}
+u_{0}
\left\{u_{1}^{*},u_{0}\right\}u_{0}^{*}\right]^{D}\sigma_{3}
\,.
\end{multline*}
The relation \eqref{eq.poissdiag} yields
\begin{multline}
\label{eq.23}
\frac{1}{2i}\left[u_{0}\left\{u_{0}^{*}\mathcal{V},u_{1}\right\}u_{0}^{*}
+u_{0}\left\{u_{1}^{*}\mathcal{V},u_{0}\right\}u_{0}^{*}\right]
=
-\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})
\\
+\mathcal{V}\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{E_{+}-E_{-}}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
\end{multline}
\item The term
$\frac{1}{2i}\left[u_{0}\left\{u_{1}^{*},H\right\}\right]^{D}$
is the sum of two terms
$$
\frac{1}{2i}\left[u_{0}\left\{u_{1}^{*},f_{\varepsilon}\right\}\right]^{D}
+
\frac{1}{2i}\left[u_{0}\left\{u_{1}^{*},\mathcal{V}\right\}\right]^{D}\,.
$$
Since
$u_{0}\partial_{p_{\ell}}u_{1}^{*}=i\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{(E_{+}-E_{-})}(\partial_{q^{k}}\Pi_{0})$
is off-diagonal while
$$
(\partial_{q^{\ell}}\mathcal{V})^{OD}=(E_{+}-E_{-})(\partial_{q^{\ell}}\Pi_{0})\,,
$$
the second term equals
$$
\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
$$
The first term a priori contains more terms because $u_{0}^{*}$ has to be differentiated:
\begin{multline*}
\frac{1}{2}
\Big[
-(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})\partial_{q^{\ell}}\left(\frac{1}{E_{+}-E_{-}}\right)(\partial_{q^{k}}\Pi_{0})
-
\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial^{2}_{q^{\ell}q^{k}}\Pi_{0})
\\
-u_{0}(\partial_{q^{\ell}}u_{0}^{*}) \frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial_{q^{k}}\Pi_{0})\Big]^{D}\,.
\end{multline*}
The first part is off-diagonal and vanishes after taking the diagonal
part.
By using again \eqref{eq.deunitOD} and \eqref{eq.dersecdiag} and $\partial_{q}\Pi_{0}=(\partial_{q}\Pi_{0})^{OD}$, we
obtain
$$
\frac{(\partial_{p_{k}}f_{\varepsilon})
(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}(\partial_{q^{k}}\Pi_{0})
(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}
+
\frac{(\partial_{p_{k}}f_{\varepsilon})
(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}(\partial_{q^{k}}\Pi_{0})\sigma_{3}
(\partial_{q^{\ell}}\Pi_{0})=0\,.
$$
Hence we have proved
\begin{equation}
\label{eq.23bis}
\frac{1}{2i}\left[u_{0}\left\{u_{1}^{*},H\right\}\right]^{D}
=\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
\end{equation}
\item The last term is
\begin{multline*}
T=\frac{1}{8}\Big[2u_{0}\left\{(\partial_{p_{k}}f_{\varepsilon})(\partial_{q^{k}}u_{0}^{*}),
u_{0}\right\}u_{0}^{*}\\
-u_{0}(\partial^{2}_{q^{k}q^{\ell}}u_{0}^{*})(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})
-
(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})(\partial^{2}_{q^{k}q^{\ell}}u_{0})u_{0}^{*}\Big]^{D}\,.
\end{multline*}
Forgetting the $\frac{1}{8}$ factor, the first part equals
$2(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})u_{0}(\partial_{q^{k}}u_{0}^{*})
(\partial_{q^{\ell}}u_{0})u_{0}^{*}$\,. By using \eqref{eq.der2u0u0}
in the second part gives
$$
8T=4(\partial_{p_{k}p_{\ell}}f_{\varepsilon})
\left[u_{0}(\partial_{q^{k}}u_{0}^{*})(\partial_{q^{\ell}}u_{0})u_{0}^{*}\right]^{D}\,.
$$
Owing to \eqref{eq.diagprod} and \eqref{eq.deunitOD}, this gives
$$
2T=
(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})
[u_{0}(\partial_{q^{k}}u_{0}^{*})]^{D}[(\partial_{q^{\ell}}u_{0})u_{0}^{*}]^{D}
+
(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
$$
With $u_{0}(\partial_{q^{k}}u_{0}^{*})=-(\partial_{q^{k}}u_{0})u_{0}^{*}$, the last term equals
\begin{equation}
\label{eq.24}
T=
-\frac{(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{2}[(\partial_{q^{k}}u_{0})u_{0}^{*}]^{D}
[(\partial_{q^{\ell}}u_{0})u_{0}^{*}]^{D}
+\frac{(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{2}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\,.
\end{equation}
\end{itemize}
By summing \eqref{eq.21},\eqref{eq.22},\eqref{eq.23},\eqref{eq.23bis},\eqref{eq.24}, we
obtain
\begin{multline*}
B_{2}^{D}= -\frac{(\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon})}{2}[(\partial_{q^{k}}u_{0})u_{0}^{*}]^{D}
[(\partial_{q^{\ell}}u_{0})u_{0}^{*}]^{D}
+ \frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}
\\
+\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}
(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})
+\varepsilon^{1+2\delta}\underline{\tilde{R}_{1}}+ \varepsilon^{1+4\delta}\tilde{R}_{1}
\,.
\end{multline*}
Hence the diagonal symbol that we seek, is
\begin{multline*}
\begin{pmatrix}
h_{+} &0\\
0&h_{-}
\end{pmatrix}
=
\begin{pmatrix}
f_{\varepsilon}(p,\tau)+E_{+}(q,\tau) & 0\\
0& f_{\varepsilon}(p,\tau)+E_{-}(q,\tau)
\end{pmatrix}
\\-
i\varepsilon
(\partial_{p_{k}}f_{\varepsilon})u_{0}^{*}\left[(\partial_{q^{k}}u_{0})u_{0}^{*}\right]^{D}u_{0}
-\varepsilon^{2}\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}u_{0}^{*}
\left[(\partial_{q^{k}}u_{0})u_{0}^{*}\right]^{D}\left[(\partial_{q^{k}}u_{0})u_{0}^{*}\right]^{D}u_{0}
\\
+\varepsilon^{2}
\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
u_{0}^{*}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})\sigma_{3}u_{0}
+
\varepsilon^{2}\frac{\partial^{2}_{p_{k}p_{\ell}}f_{\varepsilon}}{2}u_{0}^{*}(\partial_{q^{k}}\Pi_{0})(\partial_{q^{\ell}}\Pi_{0})u_{0}
\,.
\end{multline*}
The symbol
$u_{0}^{*}\left[(\partial_{q^{k}}u_{0})u_{0}^{*}\right]^{D}u_{0}$ equals
\begin{multline*}
u_{0}^{*}\Pi_{0}(\partial_{q^{k}}u_{0})u_{0}^{*}\Pi_{0}u_{0}+
u_{0}^{*}(1-\Pi_{0})(\partial_{q^{k}}u_{0})u_{0}^{*}(1-\Pi_{0})u_{0}
\\
=
P_{+}u_{0}^{*}(\partial_{q^{k}}u_{0})P_{+}
+
(1-P_{+})u_{0}^{*}(\partial_{q^{k}}u_{0})(1-P_{+})
=-i
\begin{pmatrix}
A_{k}&0\\
0& -A_{k}
\end{pmatrix}\,.
\end{multline*}
For the last term we deduce from $\Pi_{0}=u_{0}P_{+}u_{0}^{*}$ and
$(\partial_{q}u_{0})u_{0}^{*}+ u_{0}(\partial_{q}u_{0})=0$,
\begin{eqnarray*}
&&
P_{+}\left[u_{0}^{*}(\partial_{q^{k}}\Pi_{0})u_{0}\right]
\left[u_{0}^{*}(\partial_{q^{\ell}}\Pi_{0})u_{0}\right]P_{+}
\\
&&\qquad
=
P_{+}\left[u_{0}^{*}(\partial_{q^{k}}u_{0})P_{+}+
P_{+}(\partial_{q^{k}}u_{0}^{*})u_{0}\right]
\left[u_{0}^{*}(\partial_{q^{\ell}}u_{0})P_{+}+
P_{+}(\partial_{q^{\ell}}u_{0}^{*})u_{0}\right]
P_{+}
\\
&&\qquad
=
P_{+}u_{0}^{*}(\partial_{q^{k}}u_{0})P_{+}u_{0}^{*}(\partial_{q^{\ell}}u_{0})P_{+}
+
P_{+}u_{0}^{*}(\partial_{q^{k}}u_{0})P_{+}(\partial_{q^{\ell}}u_{0}^{*})u_{0}P_{+}
\\
&&\qquad\qquad
+
P_{+}(\partial_{q^{k}}u_{0}^{*})u_{0}u_{0}^{*}(\partial_{q^{\ell}}u_{0})P_{+}
+
P_{+}(\partial_{q^{k}}u_{0}^{*})u_{0}P_{+}(\partial_{q^{\ell}}u_{0}^{*})u_{0}P_{+}
\\
&&\qquad
=
-P_{+}u_{0}^{*}(\partial_{q^{k}}u_{0})(1-P_{+})u_{0}^{*}(\partial_{q^{\ell}}u_{0})P_{+}
= X_{k}\overline{X}_{\ell}\,.
\end{eqnarray*}
Taking the bracket with $(1-P_{+})$ is even simpler and gives
\begin{multline*}
(1-P_{+})\left[u_{0}^{*}(\partial_{q^{k}}\Pi_{0})u_{0}\right]
\left[u_{0}^{*}(\partial_{q^{\ell}}\Pi_{0})u_{0}\right](1-P_{+})
\\
=
-(1-P_{+})u_{0}^{*}(\partial_{q^{k}}u_{0})P_{+}u_{0}^{*}(\partial_{q^{\ell}}u_{0})(1-P_{+})
=
\overline{X_{k}}X_{\ell}\,.
\end{multline*}
This ends the proof.
{\hfill $\qed${}
\subsection{Discussion about the adiabatic approximation of the Born-Oppenheimer Hamiltonian}
\label{se.discuss}
The Theorem~\ref{th.BornOp} is a direct application of
Proposition~\ref{pr.hameff} by taking
$$
f_{\varepsilon}(p)= \varepsilon^{2\delta}\tau'\tau''|p|^{2}\gamma(\tau'\tau''|p|^{2})\,.
$$
The operators
\begin{eqnarray*}
h_{\pm, BO}(q,\varepsilon D_{q},\varepsilon)&=&
\varepsilon^{2\delta}\tau'\tau''\left[\sum_{k=1}^{d}(p_{k}
\mp \varepsilon A_{k})^{2}\right]^{Weyl}+\varepsilon^{2+2\delta}\tau'\tau''\sum_{k=1}^{d}|X_{k}|^{2}\\
&=&
\varepsilon^{2+2\delta}\tau'\tau''\left[
-\Delta
+ |A|^{2}
\mp \left[(\frac{1}{i}\nabla).A
+ A.(\frac{1}{i}\nabla)\right]
+|X|^{2}\right]
\\
&=&
\varepsilon^{2+2\delta}\tau'\tau''
\left[
|\frac{1}{i}\nabla \mp A|^{2}+|X|^{2}
\right]
\,,
\end{eqnarray*}
is nothing but
the usual adiabatic effective Hamiltonians which can be found in the
physics literature, including the Born-Huang potential
$|X|^{2}=\sum_{k=1}^{d}|X_{k}|^{2}$\,.
We refer to \cite{PST} and \cite{PST2} for a discussion of the various
presentations of the calculations and additional references.\\
Even in the region $\left\{\sqrt{\tau'\tau''}|p|\leq
r_{\gamma}\right\}$ with $p$ quantized into $\varepsilon D_{q}$,
this approximation makes sense, only for
$\delta>0$, because of the additional term
$$
\mp
2\frac{\varepsilon^{2+4\delta}\tau'\tau''}{E_{+}-E_{-}}\bar{X_{k}}X_{\ell}
(\varepsilon \partial_{q_{k}})(\varepsilon \partial_{q_{\ell}})
$$
coming from the last terms of \eqref{eq.h+} and \eqref{eq.h-}, that we
have included in the remainder. It is
not surprising (see \cite{Sor}, \cite{MaSo}) that the degree of the
differential operators increases with the degree in $\varepsilon$ in
the adiabatic expansion of Schr{\"o}dinger type Hamiltonians. The argument
of physicists says that this effective Hamiltonian is used
for relatively small frequencies (or momentum) so that $X_{k}X_{\ell}p^{k}p^{\ell}$ is negligible
w.r.t $|p-\varepsilon A|^{2}+\varepsilon^{2}|X|^{2}$\,. The introduction of
the additional factor $\varepsilon^{2\delta}$ with $\delta >0$
provides
a mathematically accurate and rather flexible implementation of this
approximation.
\section{Adaptation of the adiabatic asymptotics to the full nonlinear
minimization problem}
\label{se.minipb}
In this section, we adapt our rather general adiabatic result to our
nonlinear problem. In a first step, we give an explicit
form of Theorem~\ref{th.BornOp} in our specific framework. Those
results are effective when applied with wave functions localized in
the frequency variable, $\psi=\chi(\sqrt{\tau_{x}\tau_{y}}\varepsilon
D_{q})\psi$ for some compactly supported $\chi$\,. It could suffice if
we considered minimizing the energy among such well prepared quantum
states. We can do better by using a partition of unity in the
frequency variable, which will be combined, in the end, with the a priori
estimates coming from the complete and reduced minimization problems.
Finally an estimate of the effect of the unitary transform $\hat{U}$
on the nonlinear term is provided.
\subsection{Adiabatic approximation for the explicit Schr{\"o}dinger
Hamiltonian}
\label{se.explred}
Let us specify the result of Theorem~\ref{th.BornOp} by going back to the coordinates
$q=(q',q'')=(x,y)$ and $\tau=(\tau',\tau'')=(\tau_{x},\tau_{y})$\,.
Provided that $V_{\varepsilon,\tau}(x,y)$
fulfills the proper assumptions, the unitary transform introduced in
Theorem~\ref{th.BornOp} transforms
$H_{Lin}$ given by (\ref{Hlin}) into the Born-Oppenheimer Hamiltonian
with a good accuracy
in the low frequency region, that is when applied to wave
functions $\psi$ such that
$\psi=\chi(\sqrt{\tau_{x}\tau_{y}}\varepsilon D_{q})\psi$ for some
compactly supported $\chi$\,.
The operator $H_{Lin}$ is the $\varepsilon$-quantization of the
symbol
\begin{eqnarray*}
&&\varepsilon^{2\delta}\tau_{x}\tau_{y}|p|^{2} + u_{0}(q,\tau)
\begin{pmatrix}
E_{+}(q,\tau,\varepsilon)& 0\\
0&E_{-}(q,\tau,\varepsilon)
\end{pmatrix}u_{0}(q,\tau)^{*}\,,\\
\text{with}&&E_{\pm}(q,\tau,\varepsilon)= V_{\varepsilon,\tau}(x,y)\pm
\Omega(\sqrt{\frac{\tau_{x}}{\tau_{y}}}x)=V_{\varepsilon,\tau}(x,y)\pm
\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}\,.
\end{eqnarray*}
The operator $u_{0}(q,\tau)$ equals
\begin{eqnarray*}
&&u_{0}(q,\tau)=
\begin{pmatrix}
C& Se^{i\varphi}\\
Se^{-i\varphi}& -C
\end{pmatrix}=u_{0}(q,\tau)^{*}\\
\text{with}&&C=\cos(\frac{\theta}{2})\,,\quad S=\sin(\frac{\theta}{2})\,.
\\
\text{and}&&\cot(\theta)=\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\quad,\quad
\varphi=\sqrt{\frac{\tau_{y}}{\tau_{x}}}y\,.
\end{eqnarray*}
\begin{proposition}
\label{pr.adiabexpl1}
Assume $\delta \in (0,\delta_{0}]$, $(\tau_{x},\tau_{y})\in (0,1]^{2}$
and assume that $V_{\varepsilon,\tau}$ belongs to $S_{u}(\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle,\frac{\frac{\tau_{x}}{\tau_{y}}dx^{2}}{\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}
x\rangle^{2}}+ \frac{\tau_{y}}{\tau_{x}}dy^{2})$ then the
matricial
potential
$$
\mathcal{V}(q,\tau,\varepsilon)=V_{\varepsilon,\tau}(x,y)+\Omega( \sqrt{\frac{\tau_{x}}{\tau_{y}}}x)
\begin{pmatrix}
\cos(\theta)&e^{i\varphi}\sin (\theta)\\
e^{-i\varphi}\sin(\theta)& -\cos(\theta)
\end{pmatrix}
$$
fulfills the assumption of Theorem~\ref{th.BornOp}. \\
Choose the function $\gamma\in \mathcal{C}^{\infty}_{0}(\field{R})$ such that $\gamma\equiv 1$ in a
neighborhood of $[-r_{\gamma}^{2},r_{\gamma}^{2}]$,
as in Theorem~\ref{th.BornOp}
and consider a cut-off function $\chi\in
\mathcal{C}_{0}^{\infty}((-r_{\gamma}^{2},r_{\gamma}^{2}))$. When
$\hat{U}=U(q,\varepsilon D_{q},\tau,\varepsilon)\in
OpS_{u}(1,g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$
is given in Theorem~\ref{th.BornOp}, the identities
\begin{eqnarray}
\nonumber
&&\hat{U}^{*}H_{Lin}\hat{U}\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})
-
\varepsilon^{2+4\delta}R_{1}(\tau,\varepsilon)=\\
\label{eq.U*HUchi}
&&
\hspace{-1.5cm}
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
{\small\begin{pmatrix}
e^{+i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}\hat{H}_{+}e^{-i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}&0
\\
0&e^{-i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}\hat{H}_{-}e^{i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}
\end{pmatrix}
}\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})
\,,\\
\nonumber
&&H_{Lin}\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})-\varepsilon^{2+4\delta}R_{2}(\tau,\varepsilon)=
\\
\label{eq.UHU*chi}
&&\hspace{-1.5cm}
\varepsilon^{2+2\delta}\tau_{x}\tau_{y} \hat{U}
{\small\begin{pmatrix}
e^{+i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}\hat{H}_{+}e^{-i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}&0\\
0&e^{-i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}\hat{H}_{-}e^{i\sqrt{\frac{\tau_{y}}{\tau_{x}}}\frac{y}{2}}
\end{pmatrix}
}\hat{U}^{*}\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})
\,,
\end{eqnarray}
hold
with
\begin{eqnarray*}
\hat{H}_{+}&=& -\partial_{x}^{2}-
\left(\partial_{y}+ i\frac{x}{2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\right)^{2}
\\
&&+ \frac{1}{\varepsilon^{2+2\delta}\tau_{x}\tau_{y}}\left[V_{\varepsilon,\tau}(x,y)+\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}\right]+W_{\tau}(x,y)
\,,\\
\hat{H}_{-}&=& -\partial_{x}^{2}-
\left(\partial_{y}-
i\frac{x}{2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\right)^{2}
\\
&&+ \frac{1}{\varepsilon^{2+2\delta}\tau_{x}\tau_{y}}\left[V_{\varepsilon,\tau}(x,y)-\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}\right]
+W_{\tau}(x,y)\,,
\\
W_{\tau}(x,y)&=&
\frac{\tau_{x}}{\tau_{y}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})^{2}}+ \frac{\tau_{y}}{\tau_{x}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})}\,,
\end{eqnarray*}
and the estimates
$$
\|R_{1}(\tau,\varepsilon)\|_{\mathcal{L}(L^{2})}+\|R_{2}(\tau,\varepsilon)\|_{\mathcal{L}(L^{2})}\leq
C
$$
which are uniform w.r.t $\delta\in (0,\delta_{0}]$, $\tau\in
(0,1]^{2}$ and $\varepsilon\in
(0,\varepsilon_{0}]$\,.
\end{proposition}
\noindent{\bf Proof: \ }{}
Following the approach of Section~\ref{se.BO}, the Hamiltonian
$H_{Lin}$ is decomposed into
\begin{eqnarray*}
&&H_{Lin}= H(q,\varepsilon D_{q},\tau,\varepsilon)+
R_{\gamma}(\varepsilon D_{q},\tau,\varepsilon)\\
\text{with}
&&H(q,p,\varepsilon)=\varepsilon^{2\delta}\tau_{x}\tau_{y}|p|^{2}\gamma(\tau_{x}\tau_{y}|p|^{2})
+\Big[u_{0}
\begin{pmatrix}
E_{+}& 0\\
0&E_{-}
\end{pmatrix}u_{0}^{*}\Big](q,\tau)\,,\\
\text{and}&&R_{\gamma}(p,\tau,\varepsilon)=\varepsilon^{2\delta}\tau_{x}\tau_{y}|p|^{2}(1-\gamma(\tau_{x}\tau_{y}|p|^{2}))\,.
\end{eqnarray*}
The Hamiltonian $H(q,\varepsilon D_{q},\tau,\varepsilon)$ fulfills the
assumptions of Theorem~\ref{th.BornOp}, with the metric
$g_{\tau}=\frac{\frac{\tau_{x}}{\tau_{y}}dx^{2}}{\langle\sqrt{\frac{\tau_{x}}{\tau_{y}}}
x\rangle^{2}}+
\frac{\tau_{y}}{\tau_{x}}dy^{2}+\frac{\tau_{x}\tau_{y}dp^{2}}{\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{2}}$, because we assumed
$$
V_{\varepsilon,\tau}\in S_{u}(\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle, \frac{\frac{\tau_{x}}{\tau_{y}}dx^{2}}{\langle\sqrt{\frac{\tau_{x}}{\tau_{y}}} x\rangle^{2}}+\frac{\tau_{y}}{\tau_{x}}dy^{2}
)
$$
while the gap $E_{+}(q,\tau,\varepsilon)-E_{-}(q,\tau,\varepsilon)$ equals
$2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}$\,.
Actually the estimates
$$
|\partial_{x}^{\alpha}\partial_{y}^{\beta}u_{0}(x,y)|\leq
C_{\alpha,\beta}\left\langle \sqrt{\frac{\tau_{x}}{\tau_{y}}}
x\right\rangle^{-|\alpha|}\left(\frac{\tau_{x}}{\tau_{y}}\right)^{\frac{|\alpha|-|\beta|}{2}}
$$
are due to
$\partial_{x}\theta=-\frac{\sqrt{\tau_{x}}}{\sqrt{\tau_{y}}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})}$ and
$\partial_{y}e^{i\varphi}=i\sqrt{\frac{\tau_{y}}{\tau_{x}}}e^{i\varphi}$\,.
Moreover the explicit computation with $\alpha+ \beta=1$ leads to
\begin{eqnarray}
\label{eq.Dxu0} &&
\begin{pmatrix}
A_{x}& X_{x}\\
\overline{X_{x}}& -A_{x}
\end{pmatrix}= iu_{0}^{*}(\partial_{x}u_{0})=
\frac{i\partial_{x}\theta}{2}
\begin{pmatrix}
0 & e^{i\varphi}\\
-e^{-i\varphi}&0
\end{pmatrix}
\\
\label{eq.Dyu0}
\text{and}
&&
\begin{pmatrix}
A_{y}& X_{y}\\
\overline{X_{y}}& -A_{y}
\end{pmatrix}= iu_{0}^{*}(\partial_{y}u_{0})=
\sqrt{\frac{\tau_{y}}{\tau_{x}}} S
\begin{pmatrix}
S & -Ce^{i\varphi}\\
-Ce^{-i\varphi} & -S
\end{pmatrix}\,.
\end{eqnarray}
Hence the effective Hamiltonians $h_{B0,\pm}(q,\varepsilon
D_{q},\varepsilon)$, when restricted to the region
$\left\{\sqrt{\tau_{x}\tau_{y}}|p|\leq r_{\gamma}\right\}$, are given by the symbols
\begin{multline*}
h_{BO,\pm}=\varepsilon^{2\delta}\tau_{x}\tau_{y}\left[p_{x}^{2} +
(p_{y}\mp \varepsilon
\sqrt{\frac{\tau_{y}}{\tau_{x}}}
\sin^{2}(\frac{\theta}{2}))^{2}
+ \frac{1}{4}(|\partial_{x}\theta|^{2}+
\frac{\tau_{y}}{\tau_{x}}\sin^{2}(\theta))
\right]\\
+ V_{\varepsilon,\tau}(x,y)\pm \sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}\,.
\end{multline*}
With
$\cos(\theta)=\frac{\sqrt{\tau_{x}}x}{\sqrt{\tau_{y}}\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}$
and
$\sin(\theta)=\frac{1}{\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}$,
the Schr{\"o}dinger Hamiltonian corresponding to the RHS is
$\varepsilon^{2+2\delta}\tau_{x}\tau_{y}e^{\pm
i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{\pm}e^{\mp i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}$
with
\begin{multline*}
\hat{H}_{\pm}=-\partial_{x}^{2}-
\left(\partial_{y}\pm i\frac{x}{2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\right)^{2}
+
\frac{\tau_{x}}{\tau_{y}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})^{2}}
+ \frac{\tau_{y}}{\tau_{x}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})}
\\
+\frac{1}{\tau_{x}\tau_{y}\varepsilon^{2+2\delta}}
\left[V_{\varepsilon,\tau}(x,y)\pm
\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}\right]\,.
\end{multline*}
The remainder term in Theorem~\ref{th.BornOp} is
$$
\varepsilon^{2+4\delta}R_{1}(q,\varepsilon
D_{q},\tau,\varepsilon)+ \varepsilon^{3+2\delta} R_{2}(q,\varepsilon
D_{q},\tau,\varepsilon)\,,
$$
with $R_{1,2}\in S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ and where
$R_{2}$ vanishes in a neighborhood of
$\left\{\sqrt{\tau_{x}\tau_{y}}|p|\leq r_{\gamma}\right\}$\,.
The first term provides the expected
$\mathcal{O}(\varepsilon^{2+4\delta})$
estimate in $L^{2}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})$\,.
\\
It remains to check the effect of truncations. All the factors,
including the left terms
$$
R_{\gamma}(p, \tau,\varepsilon)\,,\quad
\varepsilon^{2\delta}\tau_{x}\tau_{y}\left[p_{x}^{2} +
(p_{y}\mp \varepsilon
\sin^{2}(\frac{\theta}{2}))^{2}
\right](1-\gamma(\tau_{x}\tau_{y}|p|^{2}))\,,
$$
belong
to $OpS_{u}(\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{2},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,.
For any $a,b$ belonging to the class $OpS_{u}(\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{2},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$,
where $b$ vanishes in a neighborhood of
$\left\{\sqrt{\tau_{x}\tau_{y}}|p|\leq r_{\gamma}\right\}$, and two cut-off functions
$\chi_{1},\chi_{2}\in \mathcal{C}^{\infty}_{0}((-r_{\gamma}^{2},r_{\gamma}^{2}))$ such that
$\chi_{1}\prec \chi_{2}$ (see Definition~\ref{de.ordercutoff}), the
pseudo-differential calculus says
\begin{eqnarray*}
&&(1-\chi_{2}(\tau_{x}\tau_{y}|p|^{2}))\sharp^{\varepsilon}
a\sharp^{\varepsilon} \chi_{1}(\tau_{x}\tau_{y}|p|^{2})\in \mathcal{N}_{u,g_{\tau}}\,,
\\
&& b\sharp^{\varepsilon}\chi_{1}(\tau_{x}\tau_{y}|p|^{2})\in
\mathcal{N}_{u,g_{\tau}}\,,
\end{eqnarray*}
with uniform estimates of all the seminorms w.r.t $\tau\in (0,1]^{2}$
and $\delta\in (0,\delta_{0}]$\,.
Applying this with $\chi_{1}=\chi$ and various $\chi_{2}$ such that
$\chi_{1}\prec \chi_{2}\prec \gamma$, implies that the
remainder terms due to truncations are
$\mathcal{O}(\varepsilon^{N})$ elements of $\mathcal{L}(L^{2})$ for
any $N\in \nz$\,, uniformly w.r.t $\tau\in (0,1]^{2}$ and $\delta\in
(0,\delta_{0}]$\,.
Fixing $N\geq 2+4\delta_{0}$ ends the proof of \eqref{eq.U*HUchi}.\\
For \eqref{eq.UHU*chi} use \eqref{eq.U*HUchi} with a cut-off function
$\chi_{1}$ such that $\chi\prec \chi_{1}$ and conjugate with $\hat{U}$~:
\begin{multline*}
H_{Lin}\hat{U}\chi_{1}(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})\hat{U}^{*}
=
\\
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
\hat{U}
\begin{pmatrix}
e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{+}e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}&0\\
0&e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{-}e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}
\end{pmatrix}\chi_{1}(\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2})\hat{U}^{*}\\
+
\varepsilon^{2+4\delta}R_{1}'(\varepsilon)\,.
\end{multline*}
Right-composing with $\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})$
and noticing that
$$\chi_{1}(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})\hat{U}^{*}\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})-
\hat{U}^{*}\chi(\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2})=R(q,\varepsilon D_{q},\tau,\varepsilon)\,,
$$ with $R\in \mathcal{N}_{u,g_{\tau}}$ lead to
\eqref{eq.UHU*chi} like above.
{\hfill $\qed${}
\subsection{Linear energy estimates for non truncated states}
\label{se.linenest}
\begin{proposition}
\label{pr.adiabexpl2}
Assume $\delta\in (0,\delta_{0}]$, $\tau=(\tau_{x},\tau_{y})\in
(0,1]^{2}$ and that $V_{\varepsilon,\tau}$
belongs to the parametric symbol class $S_{u}(\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle,\frac{\frac{\tau_{x}}{\tau_{y}}dx^{2}}{\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle^{2}}+
\frac{\tau_{y}}{\tau_{x}}dy^{2})$\,. Set $\hat{\chi}=
\chi(\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2})$ for $\chi\in
\mathcal{C}^{\infty}_{0}((-r_{\gamma}^{2},r_{\gamma}^{2}))$\,.
When
$\hat{U}$ is the unitary semiclassical operator
$U(q,\varepsilon D_{q},\tau,\varepsilon)\in OpS_{u}(1,g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$,
given in Theorem~\ref{th.BornOp} and parametrized by a truncation
in $\left\{\sqrt{\tau_{x}\tau_{y}}|p|< r_{\gamma}\right\}$, then for any $\chi\in
\mathcal{C}^{\infty}_{0}((-r_{\gamma}^{2},r_{\gamma}^{2}))$ the estimates
\begin{eqnarray}
\label{eq.U*HU}
&&
\left| \left\langle\psi\,,\,\big[\hat{U}^{*}H_{Lin}\hat{U}
-
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
H_{BO}
\big]
\psi\right\rangle\right|
\leq
C\varepsilon^{1+2\delta}\left[\varepsilon^{1+2\delta}\|\psi\|_{L^{2}}^{2}
\right.\\
\nonumber
&&\hspace{1cm}
+\|(1-\hat{\chi})\psi\|_{L^{2}}^{2}
\left.
+
\|(1-\hat{\chi})\psi\|_{L^{2}}\times
\|\sqrt{\tau_{x}\tau_{y}}|\varepsilon D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}
\right]\,,
\\
\label{eq.UHU*}
&&
\left| \left\langle\psi\,,\,\big[H_{Lin}
-
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
\hat{U}
H_{BO}
\hat{U}^{*}\big]
\psi\right\rangle\right|
\leq
C\varepsilon^{1+2\delta}\left[\varepsilon^{1+2\delta}\|\psi\|_{L^{2}}^{2}
\right.
\\
\nonumber
&&\hspace{1cm}\left.+\|(1-\hat{\chi})\psi\|_{L^{2}}^{2} +
\|(1-\hat{\chi})\psi\|_{L^{2}}\times
\|\sqrt{\tau_{x}\tau_{y}}|\varepsilon D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}
\right]\,,
\end{eqnarray}
hold uniformly w.r.t $\tau\in (0,1]^{2}$ and $\delta\in
(0,\delta_{0}]$, for all $\psi\in L^{2}(\field{R}^{2}; \field{C}} \newcommand{\nz}{\field{N}^{2})$,
with
\begin{eqnarray*}
H_{BO}&=&
\begin{pmatrix}
e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{+}e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}&0\\
0&e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{-}
e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}
\end{pmatrix}\\
\hat{H}_{\pm}&=& -\partial_{x}^{2}-
\left(\partial_{y}\pm i\frac{x}{2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\right)^{2}
+ \varepsilon^{-2-2\delta}\left[V_{\varepsilon,\tau}(x,y)\pm \sqrt{1+x^{2}}\right]
\\
&&\hspace{6cm}+
W_{\tau}(x,y)\,,
\\
W_{\tau}(x,y)&=&
\frac{\tau_{x}}{\tau_{y}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})^{2}}+ \frac{\tau_{y}}{\tau_{x}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})}
\,.
\end{eqnarray*}
\end{proposition}
\noindent{\bf Proof: \ }
Set
$$
\mathcal{D}=\hat{U}^{*}H_{Lin}\hat{U}
-\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
\begin{pmatrix}
e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{+}e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}&0\\
0&e^{-i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}\hat{H}_{-}
e^{i\frac{\sqrt{\tau_{y}}}{2\sqrt{\tau_{x}}}y}
\end{pmatrix}
$$
and bound the terms $\langle \hat{\chi}\psi\,,\,
\mathcal{D} \psi\rangle$ and $\langle (1-\hat{\chi}
\psi\,,\, \mathcal{D} \hat{\chi}\psi\rangle$
by $C\varepsilon^{2+4\delta}\|\psi\|_{L^{2}}^{2}$ with the help of
Proposition~\ref{pr.adiabexpl1}.
The remaining term is
$$
\langle (1-\hat{\chi})\psi\,,\, \mathcal{D}
(1-\hat{\chi})\psi\rangle\,,
\quad\text{with}\quad \hat{\chi}=\chi(\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2})\,.
$$
The operator $\mathcal{D}$ can be decomposed according to
$\mathcal{D}=\varepsilon^{2+2\delta}\tau_{x}\tau_{y}\mathcal{D}_{kin}+
\mathcal{D}_{pot}$ with
\begin{eqnarray}
\label{eq.Dkin}
\hspace{-0.8cm}
\mathcal{D}_{kin}&=&\hat{U}^{*}|D_{q}|^{2}\hat{U}-
\begin{pmatrix}
|D_{q}-A|^{2}+|X|^{2}&0\\
0&|D_{q}+A|^{2}+|X|^{2}
\end{pmatrix}\,,
\\
\label{eq.Dpot}
\hspace{-0.8cm}
\mathcal{D}_{pot}
&=&
\hat{U}^{*}\hat{U}_{0}
\begin{pmatrix}
E_{+}&0\\
0&E_{-}
\end{pmatrix}
\hat{U}_{0}^{*}\hat{U}
-V_{\varepsilon,\tau}(x,y)-
\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}
\begin{pmatrix}
+1&0\\
0&-1
\end{pmatrix}\,.
\end{eqnarray}
The normalization in \eqref{eq.Dpot} allows to use directly the
semiclassical calculus if one remembers that
$\hat{U}_{0}^{*}\hat{U}=\Id +\varepsilon \hat{R}$
and $\hat{U}^{*}\hat{U}_{0}=\Id +\varepsilon \hat{R}'$ with
$\varepsilon^{-2\delta}R, \varepsilon^{-2\delta}R' \in
S_{u}(\frac{1}{\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$
and
$E_{\pm}(q,\tau,\varepsilon)=V_{\varepsilon,\tau}(x,y)\pm\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}$\,.
We obtain
$$
\left|\left\langle (1-\hat{\chi})\psi\,,\,
\mathcal{D}_{pot} (1-\hat{\chi}))\psi\right\rangle\right|
\leq C \varepsilon^{1+2\delta}\|(1-\hat{\chi})\psi\|_{L^{2}}^{2}\,,
$$
which corresponds to the second term of our right-hand sides.\\
The kinetic energy term \eqref{eq.Dkin} is decomposed into
$$
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}\mathcal{D}_{kin}=\varepsilon^{2+2\delta}\tau_{x}\tau_{y}
\mathcal{D}_{kin}^{0}+
\mathcal{D}_{kin}^{1}\,,
$$
with
\begin{eqnarray}
\label{eq.Dkin1}
\mathcal{D}_{kin}^{1}&=& \varepsilon^{2\delta}\hat{U}^{*}\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2}\hat{U}-
\varepsilon^{2\delta}\hat{U}_{0}^{*}\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2}\hat{U}_{0}\,,\\
\label{eq.Dkin0}
\mathcal{D}_{kin}^{0}&=& \hat{U}_{0}^{*}
|D_{q}|^{2}\hat{U}_{0}-
\begin{pmatrix}
|D_{q}-A|^{2}+|X|^{2}&0\\
0&|D_{q}+A|^{2}+|X|^{2}
\end{pmatrix}\,.
\end{eqnarray}
Writing \eqref{eq.Dkin1} in the form
$$
\mathcal{D}_{kin}^{1}=
\varepsilon^{2\delta}(\hat{U}^{*}-\hat{U}_{0}^{*})\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2}\hat{U}+
\varepsilon^{2\delta}\hat{U}_{0}^{*}\tau_{x}\tau_{y}|\varepsilon
D_{q}|^{2}(\hat{U}-\hat{U}_{0})\,,
$$
while $\tau_{x}\tau_{y}|\varepsilon D_{q}|^{2}\in OpS_{u}(\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{2},g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, $\hat{U},\hat{U}_{0}\in
OpS_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ and
$\varepsilon^{-1-2\delta}(\hat{U}-\hat{U}_{0})\in
OpS_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$
\,, leads to
$$
\left|\left\langle (1-\hat{\chi})\psi\,,\,
\mathcal{D}_{kin}^{1}(1-\hat{\chi})
\psi\right\rangle\right|
\leq C\varepsilon^{1+4\delta}\|(1-\hat{\chi})\psi\|_{L^{2}}^{2}\,,
$$
which is even smaller than the $\mathcal{D}_{pot}$ upper bound.\\
In \eqref{eq.Dkin0}, the first term can be computed via
$$
\langle\Phi\,,\,
U_{0}^{*}|D_{q}|^{2}U_{0}(q)\Psi\rangle=\sum_{q_{j}\in \left\{x,y\right\}}
\langle
D_{q_{j}}(u_{0}(q,\tau)\Phi)\,,\, D_{q_{j}}(u_{0}(q,\tau)\Psi)\rangle\,,
$$
with $D_{q_{j}}(u_{0}f)=u_{0}(D_{q_{j}}-iu_{0}^{*}\partial_{q_{j}}u_{0})f$ and equals
\begin{eqnarray*}
\hat{U}_{0}^{*}|D_{q}|^{2}\hat{U}_{0}&=&
|D_{q}-iu_{0}^{*}\partial_{q}u_{0}|^{2}
\\
&=&
\sum_{q_{j}\in \{x,y\}}\Big[D_{q_{j}}^{2}-(iu_{0}^{*}(q)\partial_{q_{j}}u_{0}(q))^{2})D_{q_{j}}
-D_{q_{j}}(iu_{0}^{*}(q)\partial_{q_{j}}u_{0}(q))^{2})
\\
&&
\hspace{6cm}+(iu_{0}^{*}(q)\partial_{q_{j}}u_{0}(q))^{2}\Big]\,.
\end{eqnarray*}
Meanwhile expanding the entries of the second term in \eqref{eq.Dkin0}
gives
$$
|D_{q}\mp A(q)|^{2}= \sum_{q_{j}\in \left\{x,y\right\}}\left[D_{q_{j}}^{2}\mp
A_{j}(q)D_{q_{j}}\mp D_{q_{j}}A_{j}(q)
+ A_{j}(q)^{2}\right]\,.
$$
By using the expressions \eqref{eq.Dxu0} and
\eqref{eq.Dyu0} for $A, X$ and $iu_{0}^{*}\partial_{q}u_{0}$, we
obtain
\begin{eqnarray*}
&&\mathcal{D}_{kin}^{0}
=\frac{1}{2}
\begin{pmatrix}
0 & R_{-}\\
R_{+} &0
\end{pmatrix}\\
\text{with}&&
R_{\pm}=\pm i(D_{x}(\partial_{x}\theta)+(\partial_{x}\theta)
D_{x})e^{\mp i\varphi}
-\sqrt{\frac{\tau_{y}}{\tau_{x}}}\sin(\theta)(D_{y}e^{\mp
i\varphi}+e^{\mp i\varphi}D_{y})
\\
\text{and}
&&
(\partial_{x}\theta)=-\frac{\sqrt{\tau_{x}}}{\sqrt{\tau_{y}}(1+\frac{\tau_{x}}{\tau_{y}}x^{2})}
\quad,\quad
\sin(\theta)=\frac{1}{\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\,.
\end{eqnarray*}
Therefore, we obtain
\begin{equation*}
\left|\left\langle
(1-\hat{\chi})\psi\,,\,\mathcal{D}_{kin}^{0}(1-\hat{\chi})
\psi\right\rangle\right|
\leq
4\max(\sqrt{\frac{\tau_{x}}{\tau_{y}}}, \sqrt{\frac{\tau_{y}}{\tau_{x}}})\||D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}
\|(1-\hat{\chi})\psi\|_{L^{2}}
\,.
\end{equation*}
and, owing to $\tau_{x},\tau_{y}\in (0,1]$,
\begin{multline*}
\left|\left\langle
(1-\hat{\chi})\psi\,,\,\varepsilon^{2+2\delta}\tau_{x}\tau_{y}\mathcal{D}_{kin}^{0}(1-\hat{\chi})
\psi\right\rangle\right|
\\
\leq
4\varepsilon^{1+2\delta}\|\sqrt{\tau_{x}\tau_{y}}|\varepsilon D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}
\|(1-\hat{\chi})\psi\|_{L^{2}}\,.
\end{multline*}
This ends the proof of \eqref{eq.U*HU}.\\
For \eqref{eq.UHU*} it suffices to replace $\psi$ in \eqref{eq.U*HU}
by $\hat{U}^{*}\psi$ with a kinetic energy cut-off function $\chi_{1}$ such that
$\chi\prec \chi_{1}$ and then to use $(1-\hat\chi_{1})\hat{U}^{*}\hat\chi\in
Op\mathcal{N}_{u,g_{\tau}}$, with uniform seminorm estimates w.r.t
$\tau\in (0,1]^{2}$ and $\delta\in (0,\delta_{0}]$\,. The
$L^{2}$-norm of the corresponding additional error term is
$\mathcal{O}(\varepsilon^{N})$, for any $N$, and one fixes $N\geq 2+4\delta_{0}$\,.
{\hfill $\qed$
\subsection{Control of the nonlinear term}
\label{se.nonlin}
In this subsection, we estimate the effect of the operator
$\hat{U}=U(q,\varepsilon D_{q},\tau,\varepsilon)$ belonging to
$OpS_{u}(1,g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ on the
nonlinear term $\int_{\field{R}^{2}}|\psi|^{4}~dxdy$\,.
\begin{proposition}
\label{pr.L4err}
Let $\hat{U}$ be the unitary operator introduced in Theorem~\ref{th.BornOp}.
The inequalities
\begin{eqnarray}
\label{eq.L4U}
&& \int |\psi(x,y)|^{4}~dxdy \geq (1-C\varepsilon^{1+2\delta})
\int |(\hat{U}\psi)(x,y)|^{4}~dxdy\,,\\
\label{eq.L4U*}
\text{and}&&
\int |\psi(x,y)|^{4}~dxdy \geq (1-C\varepsilon^{1+2\delta})
\int |(\hat{U}^{*}\psi)(x,y)|^{4}~dxdy
\end{eqnarray}
hold for any
$\psi\in L^{4}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})$\,.
\end{proposition}
\noindent{\bf Proof: \ }
For $\psi_{1}$ and $\psi_{2}$ belonging to $L^{4}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})$,
the local relations
\begin{eqnarray*}
|\psi_{1}|^{4}(q)&=&\left(|\psi_{2}(q)|^{2}+2\Real \langle
\psi_{2}(q)\,,\,
(\psi_{1}-\psi_{2})(q)\rangle+|\psi_{1}(q)|^{2}\right)^{2}\\
&=&
|\psi_{2}(q)|^{4}+2|\psi_{2}|^{2}|\psi_{1}-\psi_{2}|^{2}
+ 4\left(\Real\langle \psi_{2}\,,\,
\psi_{1}-\psi_{2}\rangle+\frac{1}{2}|\psi_{1}-\psi_{2}|^{2}
\right)^{2}
\\
&&\hspace{6cm}
+4|\psi_{2}|^{2}\Real\langle
\psi_{2}\,,\,(\psi_{1}-\psi_{2})\rangle
\\
&\geq&
|\psi_{2}(q)|^{4}-4|\psi_{2}(q)|^{3}|\psi_{1}(q)-\psi_{2}(q)|\,,
\end{eqnarray*}
is integrated w.r.t $q=(x,y)\in \field{R}^{2}$, with H{\"o}lder inequality, into
$$
\|\psi_{1}\|_{L^{4}}^{4}=\int |\psi_{1}|^{4}~dxdy \geq \|\psi_{2}\|_{L^{4}}^{4} -
4\|\psi_{2}\|_{L^{4}}^{3}\|\psi_{1} - \psi_{2}\|_{L^{4}}\,.
$$
With $\psi_{1}=\hat{U}_{0}\psi=u_{0}(q)\psi$,
$|\psi(q)|^{2}=|\psi_{1}(q)|^{2}$ for all $q\in \field{R}^{2}$, and $\psi_{2}=
\hat{U}\psi=\psi_{1}+(\hat{U}-\hat{U}_{0})\psi$, we obtain
$$
\|\psi\|_{L^{4}}^{4}=\|\psi_{1}\|_{L^{4}}^{4}\geq
\|\hat{U}\psi\|_{L^{4}}^{4}-4\|\hat{U}\psi\|_{L^{4}}^{3}\|(\hat{U}-\hat{U_{0}})\psi\|_{L^{4}}\,.
$$
The operator $\hat{U}-\hat{U_{0}}$ equals
$\varepsilon^{1+2\delta}r(q,\varepsilon D_{q},\tau,\varepsilon)$ with
$r$ belonging to the class $S_{u}\Big(\frac{1}{\langle \sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle\langle
\sqrt{\tau_{x}\tau_{y}}p\rangle^{\infty}}, g_{\tau}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})\Big)$, where we
recall $g_{\tau}=\frac{\frac{\tau_{x}}{\tau_{y}}dx^{2}}{\langle
\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle^{2}}
+\frac{\tau_{y}}{\tau_{x}}dy^{2}+\frac{\tau_{x}\tau_{y}dp^{2}}{\langle\sqrt{\tau_{x}\tau_{y}}
p\rangle^{2}}$\,. After introducing the isometric transform on
$L^{4}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})$
$$
(T_{\varepsilon,\tau}\varphi)(x,y)=
\varepsilon\sqrt{\tau_{x}\tau_{y}}\varphi(\varepsilon \sqrt{\tau_{x}\tau_{y}} x,
\varepsilon \sqrt{\tau_{x}\tau_{y}} y)\,,
$$
the difference $\hat{U}-\hat{U}_{0}$ becomes
$$
\hat{U}-\hat{U}_{0}=\varepsilon^{1+\delta}T_{\varepsilon,\tau}^{-1}r_{1}(\varepsilon \tau_{x}x,
\varepsilon \tau_{y}y, D_{x}, D_{y},\varepsilon,\tau)T_{\varepsilon,\tau}
$$
with $r_{1}$ uniformly bounded in
$S(1, \frac{d(\tau_{x}x)^{2}}{\langle
\tau_{x}x\rangle^{2}}+d(\tau_{y}y)^{2}+\frac{dp^{2}}{\langle
p\rangle^{2}};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,. A fortiori, the symbol $r_{1}(\varepsilon \tau_{x}x,\varepsilon
\tau_{y}y,\xi,\eta;\varepsilon,\tau)$ is uniformly bounded
in
$S(1, dq^{2}+\frac{dp^{2}}{\langle p\rangle^{2}})$ and the
Lemma~\ref{le.Taylor} below provides the uniform bound
$$
\|\hat{U}-\hat{U}_{0}\|_{\mathcal{L}(L^{4})}\leq C_{0}\varepsilon^{1+\delta}\,.
$$
We have proved
$$
\|\psi\|_{L^{4}}^{4}\geq \|\hat{U}\psi\|_{L^{4}}^{4}-C\varepsilon^{1+2\delta}\|\hat{U}\psi\|_{L^{4}}^{3}\|\psi\|_{L^{4}}\,.
$$
which implies \eqref{eq.L4U}. The second inequality \eqref{eq.L4U*} is
proved similarly with $\hat{U}^{*}=\hat{U}_{0}^{*}+ (\hat{U}^{*}-\hat{U}_{0}^{*})$\,.
{\hfill $\qed$
The result below is a particular case of the general $L^{p}$ bound,
$1<p<\infty$, for pseudodifferential operator in
$OpS(1,dq^{2}+\frac{dp^{2}}{\langle p\rangle^{2}};
\mathcal{L}(\mathcal{H}_{1};\mathcal{H}_{2}))$, $\mathcal{H}_{i}$
Hilbert spaces, stated in \cite{Tay}-Proposition~5.7 and relying on
Calderon-Zygmund analysis of singular integral operators.
\begin{lemme}
\label{le.Taylor}
For any $p\in(1,+\infty)$, there exists a seminorm $\mathbf{n}$ on
$S(1,dq^{2}+\frac{dp^{2}}{\langle p\rangle^{2}};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ such that
$$
\forall a\in S(1, dq^{2}+\frac{dp^{2}}{\langle p\rangle^{2}};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N})),\quad
\|a(q,D_{q})\|_{\mathcal{L}(L^{p})}
\leq \mathbf{n}(a)\,.
$$
\end{lemme}
\noindent{\bf Proof: \ }
The Proposition~5.7 of \cite{Tay} says that for any $a\in S(1,
dq^{2}+\frac{dp^{2}}{\langle p\rangle}^{2}; \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$,
the operator $a(q,D_{q})$ is bounded on
$L^{p}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})$\,. It is not difficult to follow the
control of the constants in the previous pages of \cite{Tay} in
order to check that $\|a(q,D_{q})\|_{\mathcal{L}(L^{p})}$ is
estimated by a seminorm of $a$\,.
More efficiently, a linear mapping from
a Fr{\'e}chet space into a Banach space is continuous as soon as it is
bounded on bounded sets. Apply this argument with the result of
\cite{Tay} to
$$
S(1,dq^{2}+\frac{dp^{2}}{\langle
p\rangle^{2}};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))\ni a\mapsto a(q,D_{q})\in
\mathcal{L}(L^{p}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2}))
\,.
$$
{\hfill $\qed$
\section{Reduced minimization problems}
\label{se.explnlmin-}
In this section, we assume that the potential $V_{\varepsilon,\tau}$ satisfies
\eqref{eq.defVepsIntro}-\eqref{eq.defvIntro}.
After the first
paragraph of this section and in the rest of the paper, we
focus on the case $\tau_{y}=1$, $\tau_{x}\to 0$ (and $\varepsilon\to
0$)\,.
Two reduced problems have to be considered: 1) the one obtained as
$\varepsilon\to 0$ and $\tau_{x}$ is fixed; 2) the one derived from
the previous one as $\tau_{x}\to 0$ and which is parametrized only by
$(G,\ell_{V})$\,.
The linear part of this latter reduced problem
is a purely quadratic Schr{\"o}dinger Hamiltonian (with a constant
magnetic field), from which many a priori information can be obtained.
This section is divided into three parts. First we specify the potential
$V_{\varepsilon,\tau}$ and check our main assumptions for the general
theory.
Then we review some properties of the reduced Gross-Pitaevskii problem
parametrized by $(G,\ell_{V})$\,. Finally we make the comparison with
the reduced Gross-Pitaevskii problem
parametrized by $(G,\ell_{V},\tau_{x})$ as $\tau_{x}\to 0$ and deduce
properties which will be necessary for the study of the complete
minimization problem.
\subsection{Reduced minimization problems}
\label{se.uppbd}
\begin{lemme}
\label{le.Vepstau}
The potential $V_{\varepsilon,\tau}$ defined by
\eqref{eq.defVepsIntro}-\eqref{eq.defvIntro} belongs to the class
$S_{u}\left(\langle\sqrt{\frac{\tau_{x}}{\tau_{y}}}x\rangle,
g_{q,\tau}\right)$ with the metric
$g_{q,\tau}= \frac{\tau_{x}dx^{2}}{\tau_{y}\langle 1+\frac{\tau_{x}}{\tau_{y}}x^{2}
\rangle}+\frac{\tau_{y}}{\tau_{x}}dy^{2}$\,.
\end{lemme}
\noindent{\bf Proof: \ }{}
After the change of variable
$(x',y')=(\sqrt{\frac{\tau_{x}}{\tau_{y}}}x,\sqrt{\frac{\tau_{y}}{\tau_{x}}}y)$,
it is equivalent to check
$$
\frac{\varepsilon^{2+2\delta}}{\ell_{V}^{2}}v(\tau_{y}x,\tau_{x}y)
+
\sqrt{1+x^{2}}-\varepsilon^{2+2\delta}
\left[
\frac{\tau_{x}^{2}}{(1+x^{2})^{2}}+
\frac{\tau_{y}^{2}}{1+x^{2}}
\right]\;\in S_{u}(\langle x\rangle, \frac{dx^{2}}{\langle
x\rangle^{2}}+dy^{2})\,.
$$
It is done if
$v(\tau_{y}x,\tau_{x}y) \in S_{u}(\langle x\rangle, \frac{dx^{2}}{\langle
x\rangle^{2}}+dy^{2})$\,.
We know $v\in S(1,\frac{dx^{2}+dy^{2}}{1+x^{2}+y^{2}})$\,.
Hence for all $(\alpha,\beta)\in \nz^{2}$ there
exists $C_{\alpha,\beta}>0$ such that
\begin{eqnarray*}
\forall \tau\in (0,1]^{2}, \forall x,y\in\field{R}^{2},\quad
|\partial_{x}^{\alpha}\partial_{y}^{\beta}\left(v(\tau_{y}x,\tau_{x}y)\right)|
&&\leq
C_{\alpha,\beta}\frac{\tau_{x}^{\beta}\tau_{y}^{\alpha}}{(1+\tau_{y}^{2}x^{2}+\tau_{x}^{2}y^{2})^{-\frac{\alpha+\beta}{2}}}\\
&&\hspace{-2cm}\leq
C_{\alpha,\beta}\frac{1}{(\frac{1}{\tau_{y}^{2}}+x^{2})^{\alpha/2}}\leq
C_{\alpha,\beta}\langle x\rangle^{1-|\alpha|}\,,
\end{eqnarray*}
which is what we seek.
{\hfill $\qed${}
If the error terms of Proposition~\ref{pr.adiabexpl2} and
Proposition~\ref{pr.L4err}
are assumed to be negligible, the energy
$\mathcal{E}_{\varepsilon}(\psi)$ of a state $\psi=\hat{U}^{*}
\begin{pmatrix}
0\\alpha_{-}
\end{pmatrix}
$ is close to
\begin{eqnarray}
\nonumber
&&\varepsilon^{2+2\delta}\tau_{x}\tau_{y}\langle a_{-}\,, \hat{H}_{-}
a_{-}\rangle + \frac{G_{\varepsilon,\tau}}{2}\int |a_{-}|^{4}~dxdy
=
\varepsilon^{2+2\delta}\tau_{x}\tau_{y}\mathcal{E}_{\tau}(a_{-})\,,\\
\text{with}
&&
\label{eq.defER}
\mathcal{E}_{\tau}(a_{-})= \langle a_{-}\,, \hat{H}_{-}
a_{-}\rangle + \frac{G}{2}\int |a_{-}|^{4}~dxdy\,,\\
\nonumber
&&
\hat{H}_{-}=
-\partial_{x}^{2}-
\left(\partial_{y}- i\frac{x}{2\sqrt{1+\frac{\tau_{x}}{\tau_{y}}x^{2}}}\right)^{2}
+ \frac{1}{\ell_{V}^{2}\tau_{x}\tau_{y}}v(\sqrt{\tau_{x}\tau_{y}}x,\sqrt{\tau_{x}\tau_{y}}y)\,.
\end{eqnarray}
with the potential
$v$ chosen from (\ref{eq.defvIntro}).\\
When $\frac{\tau_{x}}{\tau_{y}}$ and $\tau_{x}\tau_{y}$ are small, in
particular in the regime $\tau_{x}\ll 1$ and $\tau_{y}=1$ that we
shall consider,
this energy is well approximated by
\begin{equation}
\label{eq.defEH}
\mathcal{E}_{H}(\varphi)=
\langle
\varphi\,,\,
\left[-\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\right]\varphi\rangle
+\frac{G}{2}\int|\varphi|^{4}\,,
\end{equation}
as this will be checked and specified in the next paragraph.
Although more general asymptotics could be considered, we concentrate
from now on the regime $\tau_{x}\ll 1$, $\tau_{y}=1$\,. The parameters
$\ell_{V}$ and $G$ are assumed to be fixed as $\tau_{x}\to 0$\,.
In order to prove that the ground states of ${\mathcal E}_{H}$ and ${\mathcal E}_{\varepsilon}$
are close, we need good estimates on the energy ${\mathcal E}_{H}$.
\subsection{Properties of the harmonic approximation}
\label{se.harmapp}
The energy functional ${\mathcal
E}_{H}$ does not any more depend on $\tau_{x}$ and is
parametrized only by $(G,\ell_{V})$\,. Let us start with its properties.
We introduce the spaces ${\mathcal H}_{1}$ and ${\mathcal H}_{2}$ which are
given by
\begin{equation}
\label{eq.defcalHs}
{\mathcal H}_{s}=\left\{u\in L^{2}(\field{R}^{2})\,
\sum_{|\alpha|+|\beta|\leq s} \|q^{\alpha}D_{q}^{\beta}u\|_{L^{2}}
<+\infty\right\}\,,\quad s=1,2, (q=(x,y))
\end{equation}
endowed with the norm $\|u\|_{{\mathcal
H}_{s}}^{2}=\sum_{|\alpha|+|\beta|\leq s}
\|q^{\alpha}D_{q}^{\beta}u\|_{L^{2}}^{2}$\,.
For a compact set $K$ of ${\mathcal H}_{s}$ and for $u\in {\mathcal H}_{s}$,
the distance $d_{s}(u,K)$ follows the usual definition $\min_{v\in K}\|u-v\|_{{\mathcal H}_{s}}$\,.
The self-ajoint operator associated with the linear part of ${\mathcal
E}_{H}$ is denoted by
$$
H_{\ell_{V}}= -\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\,, \quad (0<\ell_{V}<+\infty)\,.
$$
Its domain is ${\mathcal H}_{2}$ while its form domain is ${\mathcal
H}_{1}$\,. Note also the compact embeddings ${\mathcal
H}_{2}\subset\subset {\mathcal H}_{1}\subset\subset L^{2}\cap L^{4}$\,.
Following the general scheme presented in \cite{HiPr,Sjo},
its spectrum equals
$$
\sigma(H_{\ell_{V}})=\left\{(1+2n_{+})r_{+}+(1+2n_{-})r_{-}\,,\quad
(n_{+},n_{-})\in \nz^{2}\right\}
$$
with $r_{\pm}=\frac{1}{2\sqrt{2}}\sqrt{1+\frac{8}{\ell_{V}^{2}}\pm\sqrt{1+\frac{4}{\ell_{V}^{2}}}}$\,.
\begin{proposition}
\label{pr.harmapp}
The functional ${\mathcal E}_{H}$ admits minima on $\left\{u\in {\mathcal
H}_{1}, \|u\|_{L^{2}}=1\right\}$\,, with a minimum value ${\mathcal
E}_{H,min}$ satisfying
$$
{\mathcal E}_{H,min}\geq r_{+}+r_{-}\geq \frac{\sqrt{2}}{2}\,.
$$
The set of minimizers $\mathrm{Argmin~}{\mathcal E}_{H}$ is a bounded subset of
${\mathcal H}_{2}$ and therefore a compact subset of $\left\{u\in {\mathcal
H}_{1}, \|u\|_{L^{2}}=1\right\}$\,. Moreover for any $\varphi\in
\mathrm{Argmin~}{\mathcal E}_{H}$, $\varphi$ is an eigenvector of $H_{0}+
G|\varphi|^{2}$\,.
Finally there exist two constants $C=C_{\ell_{V},G}>0$ and
$\nu=\nu_{\ell_{V},G}\in (0,1/2]$ such that the conditions $u\in {\cal
H}_{1}$,
$\|u\|_{L^{2}}=1$ and ${\mathcal E}_{H}(u)\leq {\mathcal E}_{H,min}+1$, imply
\begin{equation}
\label{eq.lojas}
d_{{\mathcal H}_{1}}(u,\mathrm{Argmin}~{\mathcal E}_{H})\leq C({\mathcal E}_{H}(u)-{\mathcal E}_{H,min})^{\nu}\,.
\end{equation}
\end{proposition}
\noindent{\bf Proof: \ }{}
On ${\mathcal H}_{1}$ and ${\mathcal H}_{2}$, the scalar products
\begin{eqnarray*}
&&\langle u\,,\, v\rangle_{1,\ell_{V}}=\langle u\,,\,
H_{\ell_{V}}v\rangle_{L^{2}}
\\
&&
\langle u\,,\, v\rangle_{2,\ell_{V}}=\langle H_{\ell_{V}}u\,,\,
H_{\ell_{V}}v\rangle_{L^{2}}
\end{eqnarray*}
provide norms $\|u\|_{k,\ell_{V}}$, $k=1,2$, respectively equivalent to $\|u\|_{{\mathcal
H}_{k}}$\,. In this proof, all the ``uniform'' estimates are
actually parametrized by $(G,\ell_{V})$\,.
The nonlinearity $\frac{G}{2}\int|u|^{4}(x)~dx$ as well as the
constraint $\|u\|_{L^{2}}=1$ are continuous functions on $L^{2}\cap
L^{4}$ while the quadratic part of ${\mathcal E}_{H}(u)$ is simply
$\|u\|_{1,\ell_{V}}^{2}$ with ${\mathcal E}_{H}(u)\geq
\|u\|_{1,\ell_{V}}^{2}\geq r_{+}+r_{-}\geq \frac{\sqrt{2}}{2}$\,.
The compact embedding ${\mathcal H}_{1}\subset\subset L^{2}\cap L^{4}$
thus implies that the infimum $\inf_{u\in {\mathcal H}_{1}\,,\,
\|u\|_{L^{2}}=1}{\mathcal E}_{H}(u)$ is achieved.\\
A minimizer $\varphi\in \mathrm{Argmin}~{\mathcal E}_{H}$ solves in a
distributional sense the Euler-Lagrange equation
$$
H_{\ell_{V}}\varphi +G|\varphi|^{2}u=\lambda_{\varphi}\varphi
$$
where $\lambda_{\varphi}$ is the Lagrange multiplier associated with the
constraint $\|\varphi\|_{L^{2}}=1$\,. By taking the scalar product with $\varphi$,
one obtains the bounds for $\lambda_{\varphi}$:
$$
{\mathcal E}_{H,min}\leq \lambda_{\varphi}\leq 2{\mathcal E}_{H,min}\,.
$$
Since ${\mathcal H}_{1}$ is also (compactly) embedded in $L^{6}(\field{R}^{2})$,
the equation
$$
H_{\ell_{V}}\varphi = -G|\varphi|^{2}\varphi+\lambda_{\varphi}\varphi
$$
ensures that $\|\varphi\|_{2,\ell_{V}}$ is uniformly bounded on
$\mathrm{Argmin}~{\mathcal E}_{H}$\,.
Therefore $\mathrm{Argmin}~{\mathcal E}_{H}$ is a bounded subset of ${\mathcal
H}_{2}$ and a compact subset of ${\mathcal H}_{1}$\,.
In an ${\mathcal H}_{1}$-neighborhood of $\varphi\in \mathrm{Argmin}~{\mathcal
E}_{H}$ ($\|\varphi\|_{L^{2}}=1$), the $L^{2}$-sphere $\left\{u\in
{\mathcal H}_{1},\, \|u\|_{L^{2}}=1\right\}$ can be parametrized by
$$
u=(1-\|v\|_{L^{2}}^{2})\varphi + v\,,\quad \langle \varphi\,,\, v\rangle_{L^{2}}=0\,.
$$
Notice also that the potential
$G|\varphi|^{2}$ is a relatively compact perturbation of
$H_{\ell_{V}}$, so that $H_{\ell_{V}}+ G|\varphi|^{2}$ is a
self-adjoint operator in $L^{2}(\field{R}^{2})$ with domain ${\mathcal H}_{2}$
and with a compact resolvent.
With $\langle \varphi\,,\,
v\rangle_{1,\ell_{V}}=-G\int_{\field{R}^{2}}|\varphi|^{2}\overline{\varphi}v~dx$
for $\varphi$ is
an eigenvector of $H_{\ell_{V}}+G|\varphi|^{2}$ and $v\perp \varphi$,
the energy ${\mathcal E}_{H}(u)$ becomes
\begin{eqnarray*}
{\mathcal E}_{H}((1-\|v\|_{L^{2}}^{2})\varphi+v)&=&\|v\|_{1,\ell_{V}}^{2} +(1-\|v\|_{L^{2}})\langle
\varphi\,,\,\varphi\rangle_{1,\ell_{V}}
\\
&&\hspace{-3cm}
-2G(1-\|v\|_{L^{2}}^{2})\Real\int_{\field{R}^{2}}|\varphi|^{2}\overline{\varphi}v~dx
+\frac{G}{2}\int_{\field{R}^{2}}|(1-\|\varphi\|_{L^{2}}^{2})\varphi+v|^{4}~dx\\
&=&\|v\|_{1,\ell_{V}}^{2}+ F_{\varphi}(v)\,,
\end{eqnarray*}
where $v$ lies in the closed subset ${\mathcal H}_{1,\varphi}=\left\{v\in {\mathcal H}_{1}, \langle
\varphi\,, v\rangle_{L^{2}}=0\right\}$ of ${\mathcal H}_{1}$ and
$F_{\varphi}(v)$ is the composition of the compact embedding ${\mathcal
H}_{1}\to L^{2}\cap L^{4}$ with a real analytic, real-valued,
functional on $L^{2}\cap L^{4}$\,. Hence on ${\mathcal H}_{1,\varphi}$ endowed
with the scalar product $\langle~,~\rangle_{1,\ell_{V}}$, the Hessian
of ${\mathcal E}_{H}((1-\|v\|_{L^{2}}^{2})\varphi+v)$ equals $\Id+
D^{2}F_{\ell_{V}}(0)$, with $D^{2}F_{\ell_{V}}(0)$ compact (and
self-adjoint)\,.
We can apply the Lojasiewicz-Simon inequality which says that there
exist two constants $C_{\varphi}>0$, $\nu_{\varphi}\in (0,1/2]$, such that
$$
\|v\|_{1,\ell_{V}}\leq C_{\varphi}\left({\mathcal E}_{H}((1-\|v\|_{L^{2}}^{2})\varphi+v)-{\mathcal E}_{H,min}\right)^{\nu_{\varphi}}\,.
$$
Since the set $\mathrm{Argmin}~{\mathcal E}_{H}$ is a compact subset of
${\mathcal H}_{1}$, it can be covered by a finite number of neighborhoods
of $\varphi_{i}\in \mathrm{Argmin}~{\mathcal E}_{H}$, $1\leq i\leq N$, where a
Lojasiewicz-Simon inequality holds. Take
$$
\nu_{\ell_{V},G}=\min_{1\leq i\leq
N}\nu_{\varphi_{i}}\quad\text{and}\quad
C_{\ell_{V},G}=2\max_{1\leq i\leq N}C_{\varphi_{i}}\,.
$$
{\hfill $\qed${}
\begin{remarque}
\label{re.loja}
The Lojasiewicz inequality is a classical result of real algebraic
geometry (see a.e. \cite{Loj,BCR}) proved by Lojasiewicz
after Tarski-Seidenberg Theorem. It is usually written as
$|\nabla f(x)|\leq C |f(x)|^{\nu}$ with $\nu\in (0,1]$ for a real
analytic function of $x$ lying around $x_{0}$ with
$f(x_{0})=0$\,. The variational form is a variant of it. It was
extended to the infinite dimensional case with applications to PDE's
by L.~Simon in \cite{Sim}.
We refer the reader also to \cite{Chi,HaJe,Hua} and
\cite{BDLM}
for recent
texts and references concerned with the infinite dimensional case or
the extension with $o$-minimal structures.
\item The nonlinear Euler-Lagrange equation is usually studied after
linearization via the Liapunov-Schmidt process.
Here using some coordinate representation of the
constraint submanifold, especially when it is a sphere for a simple
norm, allows to use directly the standard result for the
minimization of real analytic functionals.
\item When the minimization problem is non degenerate at every
$\varphi\in \mathrm{Argmin}~{\mathcal E}_{H}$, i.e. in the present
case when the kernel of $\Id+D^{2}F_{\varphi}(0)$ is restricted to
$\left\{0\right\}$, the compact set $\mathrm{Argmin}~{\mathcal E}_{H}$ is
made of a finite number of point. When $\ell_{V}$ is fixed so that
$r_{+}$ and $r_{-}$ are rationally independent, the spectrum of
$H_{\ell_{V}}$ is made of simple eigenvalues and when $G$ is small enough,
$G < G_{\ell_{V}}$, the non degeneracy assumption is satisfied via a perturbation
argument from the case $G=0$\,. For large $G$, we can only say that
the set of $(G,\ell_{V})\in (0,+\infty)^{2}$ such that all the
minima are non degenerate, $\nu_{\ell_{V},G}=1/2$, is a subanalytic
subset of $\field{R}^{2}$\,. In our case with a linear part
$H_{\ell_{V}}$ which is a complex operator with no rotational
symmetry, no standard methods like in \cite{AJR} allow to reduce the
minimization problem to some radial nonlinear ODE.
From the
information given by the Lowest-Landau-Level reduction, when $G$ and
$\ell_{V}$ are large, the supposed hexagonal symmetry, after
removing some trivial rotational invariance, of the
problem (see \cite{ABN,Nie}) suggests that there are
presumably several minimizers.
\end{remarque}
A change of variable $\varphi(x,y)e^{-ixy/4}=\alpha u(\alpha x,\alpha y)$ with $\alpha^2=1/(\ell_{V}\sqrt{G})$
leads to \begin{equation}{\mathcal E}_H(\varphi)=\frac 1{\ell_{V}\sqrt{G}}\tilde { \mathcal E}_H(u)=\frac 1{\ell_{V}\sqrt{G}}\int
|(\nabla - \frac i 2 \ell_{V}\sqrt{G} e_{z}\times \vec{r})
u|^2+G(r^2|u|^2+\frac 12 |u|^4)\,,
\end{equation}
with the notations $\vec{r}=
\begin{pmatrix}
x\\y
\end{pmatrix}
$ $r=|\vec{r}|$\,.
This implies that $\ell_{V}\sqrt{G}$ is equivalent to a rotation value. We have the following results from
the literature\begin{itemize}\item when $\ell_{V}\sqrt{G}$ is small and $G$ is large, the minimizer is unique
up to rotation and vortex-free \cite{AJR}: namely $u(x,y)=f(r)e^{ic}$ for some real number $c$, where $f$ does not vanish.
If $\ell_{V}\sqrt{G}=0$, this is an adaptation of a result of \cite{BrOs}. When $\ell_{V}\sqrt{G}$ is non zero, this
requires refined estimates for the jacobian.
\item when $\ell_{V}\sqrt{G}$ is large, then vortices are expected in the system and this can be
analyzed in details in the LLL regime (lowest Landau level) if additionally $\sqrt G/\ell_{V}$ is small \cite{AB,ABN}.
More precisely, if $\sqrt G/\ell_{V}$ is small, then
\begin{equation} \inf \tilde {\mathcal E}_H -\frac 12- \inf E_{LLL}=o \left (\frac { \sqrt G}{\ell_{V}}\right )\end{equation}
where \begin{equation}E_{LLL}(u)=\int G(r^2|u|^2+\frac 12 |u|^4)\end{equation} for functions $u$ such that $u(x,y)e^{\ell_{V}\sqrt{G}r^2/4}$
is a holomorphic function of $x+iy$. This space is called the LLL. If $u$ is a ground state of $\tilde {\mathcal E}_H$
and $w$ its projection onto the LLL, then $|u-w|$ tends to 0 in $H^1$ and $C^{0,\alpha}$ as $\sqrt G/\ell_{V}$
tends to 0. If additionally, $\ell_{V}\sqrt{G}$ is large, then one can estimate $\inf E_{LLL}$ \cite{ABN} thanks to test
functions with vortices and $\inf E_{LLL}=O\left (\frac { \sqrt G}{\ell_{V}}\right )$.\item if $\ell_{V}\sqrt{G}$ is large,
and $\sqrt G/\ell_{V}$ is large, then this is a Thomas Fermi regime where the energy can be estimated as well \cite{Aft}
and is of order $\sqrt G/\ell_{V}$.
\end{itemize}
We complete the previous result with another comparison statement
which will be useful in the sequel.
\begin{proposition}
\label{pr.harmH2}
There exists $C=C_{\ell_{V},G}>0$ such that when
$u\in \mathcal{H}_{1}$ satisfy $\mathcal{E}_{H}(u)\leq
\mathcal{E}_{H,min}+1$, $\|u\|_{L^{2}}=1$, and solves
$$
H_{\ell_{V},G}u+G|u|^{2}u= \lambda_{u} u + r
$$
with $\lambda_{u}\in \field{R}$ and $r\in L^{2}$, then
\begin{itemize}
\item $u\in \mathcal{H}_{2}$\,;
\item there exists $u_{0}\in \mathrm{Argmin}~\mathcal{E}_{H}$, with
Lagrange multiplier $\lambda_{u_{0}}$, such that
$$
|\lambda_{u}-\lambda_{u_{0}}|+\|u-u_{0}\|_{\mathcal{H}_{2}}\leq
C\left(\|r\|_{L^{2}}+ \left(\mathcal{E}_{H}(u)-\mathcal{E}_{H,min}\right)^{\nu}\right)\,,
$$
where $\nu=\nu_{\ell_{V},G}\in (0,\frac{1}{2}]$ is the exponent given
in Proposition~\ref{pr.harmapp}.
\end{itemize}
\end{proposition}
\noindent{\bf Proof: \ }{}
Since $\mathrm{Argmin}~\mathcal{E}_{H}$ is compact,
Proposition~\ref{pr.harmapp} already provides $u_{0}\in
\mathrm{Argmin}~\mathcal{E}_{H}$ such that
$$
\|u-u_{0}\|_{\mathcal{H}_{1}}\leq
C(\mathcal{E}_{H}(u)-\mathcal{E}_{H,min})^{\nu}\,.
$$
Taking the difference of the equation for $u$ and the Euler-Lagrange
equation for $u_{0}$, we obtain
$$
H_{\ell_{V},G}(u-u_{0})=\left(\lambda_{u}-\lambda_{u_{0}}\right)u_{0}+
\lambda_{u}(u-u_{0})+G(|u_{0}|^{2}u_{0}-|u|^{2}u)+r\,.
$$
Taking the scalar product with $u_{0}$, with
$$
\||u_{0}|^{2}u_{0}-|u|^{2}u\|_{L^{2}}\leq
C\left(\mathcal{E}_{H}(u_{0})+\mathcal{E}_{H}(u)\right)
\|u-u_{0}\|_{\mathcal{H}_{1}}\leq C'\left(\mathcal{E}_{H}(u)-\mathcal{E}_{H,min}\right)^{\nu}
$$
implies
$$
|\lambda_{u}-\lambda_{u_{0}}|\leq C'' \left(\|r\|_{L^{2}}+ \left(\mathcal{E}_{H}(u)-\mathcal{E}_{H,min}\right)^{\nu}\right)\,.
$$
Using the ellipticity of $H_{\ell_{V},G}$ and the equivalence of the
norms $\|\varphi\|_{\mathcal{H}_{2}}$ and $\|H_{\ell_{V},G}\varphi\|_{L^{2}}$ ends
the proof.
{\hfill $\qed${}
\subsection{Comparison of the two reduced minimization problems}
\label{se.compred}
In the regime $\tau_{y}=1$ and $\tau_{x}\to 0$, while $\ell_{V}>0$ and
$G>0$ are fixed, we compare the two minimization problems for the energies ${\mathcal
E}_{\tau}$ and ${\mathcal E}_{H}$ defined in
\eqref{eq.defER}-\eqref{eq.defEH}.
We start with the next Lemma which is a simple application of the so
called IMS localization formula (see a.e. \cite{CFKS}). We shall use
the functional spaces ${\mathcal H}_{s}$ defined by \eqref{eq.defcalHs}
associated with $\mathcal{E}_{H}$ as well as the standard Sobolev
spaces $H^{s}(\field{R}^{2})$ associated with $\mathcal{E}_{\tau}$, with
$s=1,2$ and $\mathcal{H}_{s}\subset H^{s}(\field{R}^{2})$\,.
\begin{lemme}
\label{le.compred}
Let $\chi_{1}, \chi_{2} \in \mathcal{C}_{b}^{\infty}(\field{R}^{2})$ satisfy
$\chi_{1}^{2}+\chi_{2}^{2}=1$, $\supp \chi_{1}\subset
\left\{x^{2}+y^{2}< 1\right\}$ and take $\alpha\in (0,\frac{1}{2}]$\,. Then the
following identity
\begin{multline}
\label{eq.IMSH}
\mathcal{E}_{H}(u)=
\mathcal{E}_{H}(\chi_{1}(\tau_{x}^{\alpha}.)u)+\mathcal{E}_{H}(\chi_{2}(\tau_{x}^{\alpha}.)u)
-
\tau_{x}^{2\alpha}\sum_{j=1}^{2}\int_{\field{R}^{2}}|(\nabla\chi_{j})(\tau_{x}^{\alpha}.)|^{2}|u|^{2}
\\
+G\int_{\field{R}^{2}}(\chi_{1}^{2}\chi_{2}^{2})(\tau_{x}^{\alpha}.)|u|^{4}\,,
\end{multline}
holds for all $u\in \mathcal{H}_{1}$,
with the same formula for $\mathcal{E}_{\tau}(u)$ when $u\in
H^{1}(\field{R}^{2})$\,. Moreover, $\mathcal{E}_{\tau}$ and $\mathcal{E}_{H}$ satisfy
\begin{multline}
\label{eq.IMSR}
\mathcal{E}_{\tau}(u)=\mathcal{E}_{H}(\chi_{1}(\tau_{x}^{\alpha}.)u)+\mathcal{E}_{\tau}(\chi_{2}(\tau_{x}^{\alpha}.)u)
- \tau_{x}^{2\alpha}\sum_{j=1}^{2}\int_{\field{R}^{2}}|(\nabla
\chi_{j})(\tau_{x}^{\alpha}.)|^{2}|u|^{2}\\
+G\int_{\field{R}^{2}}(\chi_{1}^{2}\chi_{2}^{2})(\tau_{x}^{\alpha}.)|u|^{4}+ R(u)\,,
\end{multline}
for all $u\in H^{1}(\field{R}^{2})$ with
$$
|R(u)|\leq\frac{1}{4}
\left(\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{\alpha}.)u)^{1/2}+
\mathcal{E}_{H}(\chi_{1}(\tau_{x}^{\alpha}.)u)^{1/2}\right)\|u\|_{L^{2}}
\tau_{x}^{1-3\alpha}\,.
$$
\end{lemme}
\noindent{\bf Proof: \ }{}
The first identity is a direct application of the IMS localization
formula (see a.e. \cite{CFKS}) which comes from the identity
$$
P\chi^{2}P-\chi P^{2}
\chi=\left[P,\chi\right]^{2}-\frac{1}{2}\left[\chi^{2},P\right]P
-\frac{1}{2}P\left[P,\chi^{2}\right]
$$
when $P$ is a differential operator of order $\leq 1$ and $\chi$ is a
$\mathcal{C}^{\infty}$ function. Simply combine it with the identity
$$
|u|^{4}=
|\chi_{1}(\tau_{x}^{\alpha}.)u|^{4}+|\chi_{2}(\tau_{x}^{\alpha}.)u|^{4}+
2\chi_{1}^{2}\chi_{2}^{2}(\tau_{x}^{\alpha}.)|u|^{4}\,.
$$
Using the same argument for $\mathcal{E}_{\tau}$ provides the same
identity after replacing $\mathcal{E}_{H}$ with $\mathcal{E}_{\tau}$, and
it suffices to compare
$\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{\alpha}.)u)$ with $\mathcal{E}_{H}
(\chi_{1}(\tau_{x}^{\alpha}.)u)$\,. The definition \eqref{eq.defvIntro} of
the potential $v$ and the condition $\alpha\leq \frac{1}{2}$
imply
$$
v(\tau_{x}^{1/2}.)= \tau_{x}(x^{2}+y^{2})\,\quad\text{on}~\supp \chi_{1}(\tau_{x}^{\alpha}.)\,.
$$
Therefore, we obtain, by setting $u_{\tau}=
\chi_{1}(\tau_{x}^{\alpha}.)u$\,,
\begin{eqnarray*}
&& |\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{\alpha}.)u)-
\mathcal{E}_{H}(\chi_{1}(\tau_{x}^{\alpha}.)u)|
=
\\
&&\qquad
\left|
\int_{\field{R}^{2}}|(\partial_{y}-i\frac{x}{2\sqrt{1+\tau_{x}x^{2}}})u_{\tau}|^{2}
- |(\partial_{y}-i\frac{x}{2})u_{\tau}|^{2}\right|\\
&&\leq
\left(
\|(\partial_{y}-i\frac{x}{2\sqrt{1+\tau_{x}x^{2}}})u_{\tau}\|_{L^{2}}+
\|(\partial_{y}-i\frac{x}{2})u_{\tau}\|_{L^{2}}
\right)
\\
&&\hspace{4cm}\times
\|\frac{\tau_{x}x^{3}/2}{1+\sqrt{1+\tau_{x}x^{2}}}\chi_{1}(\tau_{x}^{\alpha}.)u\|_{L^{2}}
\\
&&
\leq\frac{1}{4}
\left(\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{\alpha}.)u)^{1/2}+
\mathcal{E}_{H}(\chi_{1}(\tau_{x}^{\alpha}.)u)^{1/2}\right)\|u\|_{L^{2}}
\tau_{x}^{1-3\alpha}\,.
\end{eqnarray*}
{\hfill $\qed${}
\begin{proposition}
\label{pr.redHR}
For any given $(\ell_{V},G)\in (0,+\infty)^{2}$, there exists
$\tau_{\ell_{V},G}>0$ such that the following properties hold when
$\tau_{x}\leq \tau_{\ell_{V},G}$\,.\\
\begin{itemize}
\item The minimization problem
$$
\inf_{u\in H^{1}(\field{R}^{^{2}})\,,\, \|u\|_{L^{2}}=1}\mathcal{E}_{\tau}(u)
$$
admits a solution $u\in H^{1}(\field{R}^{2})$\,.
\item A solution $u\in H^{1}(\field{R}^{2})$ to the above minimization
problem,
solves an Euler-Lagrange equation
$$
\left[-\partial_{x}^{2}-(\partial_{y}-\frac{i}{2}\frac{x}{\sqrt{1+\tau_{x}x^{2}}})^{2}
+ \frac{v(\tau_{x}^{1/2}.)}{\ell_{V}^{2}\tau_{x}}+G|u|^{2}\right]u=\lambda_{u}u
$$
with $0\leq \lambda_{u}\leq 2\mathcal{E}_{\tau,min}$ and belongs to $H^{2}(\field{R}^{2})$\,.
\item Moreover the minimum value
$\mathcal{E}_{\tau,min}=\min_{u\in H^{1}(\field{R}^{2})\,,\,
\|u\|_{L^{2}}=1}\mathcal{E}_{\tau}(u)$ satisfies the estimate
$$
|\mathcal{E}_{\tau,min}-\mathcal{E}_{H,min}|\leq
C_{\ell_{V},G}\tau_{x}^{2/3}\,.
$$
\item For $u\in \mathrm{Argmin}~\mathcal{E}_{\tau}$ and any pairs $\chi=(\chi_{1},\chi_{2})$ in $
\mathcal{C}^{\infty}_{b}(\field{R}^{2})^{2}$ such that
$\chi_{1}^{2}+\chi_{2}^{2}=1$ with $\supp
\chi_{1}\subset\left\{x^{2}+y^{2}<1\right\}$ and $\chi_{1}\equiv 1$
in $\left\{x^{2}+y^{2}\leq 1/2\right\}$, the functions
$\chi_{j}(\tau_{x}^{1/9}.)u$, $j=1,2$, satisfy
\begin{eqnarray}
\label{eq.estimu2}
&& \|\chi_{2}(\tau_{x}^{1/9}.)u\|^{2}_{L^{2}}\leq
C_{\chi,\ell_{V},G}\tau_{x}^{2/3}\\
&&
\label{eq.estimEnu1}
\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{1/9}.)u)\leq \mathcal{E}_{\tau,min}+ C_{\chi,\ell_{V},G}\tau_{x}^{2/3}
\\
\label{eq.estimH2u1}
&& d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{1/9}.)u,
\mathrm{Argmin}~\mathcal{E}_{H})\leq
C_{\chi,\ell_{V},G}\tau_{x}^{2\nu_{\ell_{V,G}}/3}\,,\\
&&
\nu_{\ell_{V},G}\in (0,\frac{1}{2}]\,.
\end{eqnarray}
\end{itemize}
A constant $C_{a,b,c}$ is a constant which is fixed once $(a,b,c)$
are given.
\end{proposition}
\noindent{\bf Proof: \ }{}
Fix $\ell_{V}$ and $G$. We drop the indices $\ell_{V},G$ in the constants. The
exponent $\alpha$ will be fixed to the value $\frac{1}{9}$ within the proof.\\
\noindent\underline{First step, upper bound for
$\inf\{\mathcal{E}_{\tau}(u)\,,\; u\in H^{1}(\field{R}^{2})\,,\,
\|u\|_{L^{2}}=1\}$:}\\
Let $\chi=(\chi_{1},\chi_{2})$
and $\tilde{\chi}=(\tilde{\chi}_{1},\tilde{\chi_{2}})$ be two pairs as
in our statement such that $\tilde{\chi}_{1}\prec\chi_{1}$ according to
Definition~\ref{de.ordercutoff}.
Take $u_{0}\in \mathrm{Argmin}~\mathcal{E}_{H}\subset
\mathcal{H}_{1}$\,. According to Proposition~\ref{pr.harmapp}, it
belongs to a bounded set
of $\mathcal{H}_{2}$ so that $\||q|^{2}u_{0}\|_{L^{2}}$, with
$q=(x,y)$, is uniformly
bounded. Hence,
$0\not\in\supp\nabla\tilde{\chi}_{j}\cup \supp
\tilde{\chi}_{1}\tilde{\chi}_{2}$ implies
$$
\int_{\field{R}^{2}}|\nabla\tilde{\chi}_{j}(\tau_{x}^{\alpha}.)|^{2}|u_{0}|^{2}=
\mathcal{O}(\tau_{x}^{4\alpha})\quad\text{while}\quad
\int_{\field{R}^{2}}\tilde{\chi}_{1}^{2}\tilde{\chi}_{2}^{2}(\tau_{x}^{\alpha}.)|u_{0}|^{4}
\geq 0\,.
$$
Lemma~\ref{le.compred} above with the pair $\tilde{\chi}$
and $\alpha\in (0,\frac{1}{2}]$ gives:
$$
\mathcal{E}_{H, min}=\mathcal{E}_{H}(u_{0})
\geq \mathcal{E}_{H}(\tilde{\chi}_{1}(\tau_{x}^{\alpha}. )u_{0})
+
\mathcal{E}_{H}(\tilde{\chi}_{2 }(\tau_{x}^{\alpha}. )u_{0})- C\tau_{x}^{6\alpha}\,.
$$
On $\supp \tilde{\chi}_{2}(\tau_{x}^{\alpha}.)$, the potential
$\frac{v(\tau_{x}^{1/2}.)}{\ell_{v}^{2}\tau_{x}}$ is bounded from below by
$\frac{1}{C'\tau_{x}^{2\alpha}}$\,. Thus we get
$$
\mathcal{E}_{H,min}\left(\|\tilde{\chi_{1}}u_{0}\|_{L^{2}}^{2}+\|\tilde{\chi}_{2}u_{0}\|_{L^{2}}^{2}\right)
\geq \mathcal{E}_{H,min}\|\tilde{\chi_{1}}u_{0}\|_{L^{2}}^{2}
+ \frac{1}{C'\tau_{x}^{2\alpha}}\|\tilde{\chi}_{2}u_{0}\|_{L^{2}}^{2} -C\tau_{x}^{6\alpha}\,,
$$
and finally
$$
\|\tilde{\chi}_{2}u_{0}\|_{L^{2}}^{2}\leq C''\tau_{x}^{8\alpha}\,,\quad
\|\tilde{\chi}_{1}u_{0}\|_{L^{2}}^{2}=1+\mathcal{O}(\tau_{x}^{8\alpha})\,,
$$
as soon as
$\tau_{x}<(C'\mathcal{E}_{H,min})^{-1/2\alpha}$\,.\\
The function $u_{1}=
\|\tilde{\chi}_{1}u_{0}\|_{L^{2}}^{-1}\tilde{\chi}_{1}u_{0}$ is
normalized with
$$
\mathcal{E}_{H,min}\leq \mathcal{E}_{H}(u_{1})\leq
\mathcal{E}_{H,min}+
\mathcal{O}(\tau_{x}^{6\alpha})\,,
$$
and $\chi_{1}(\tau_{x}^{\alpha}.)u_{1}=u_{1}$,
$\chi_{2}(\tau_{x}^{\alpha}.)u_{1}=0$\,.
Applying the second formula of Lemma~\ref{le.compred} with, now, the
pair $\chi$, leads to
\begin{eqnarray*}
&& \mathcal{E}_{\tau}(u_{1})= \mathcal{E}_{H}(u_{1})+ R(u_{1})
= \mathcal{E}_{H,min}+ \mathcal{O}(\tau_{x}^{6\alpha})+ R(u_{1})\\
\text{with}
&& R(u_{1})\leq \frac{1}{4}(\mathcal{E}_{\tau}(u_{1})^{1/2}+
(\mathcal{E}_{H,min}+\mathcal{O}(\tau_{x}^{6\alpha}))^{1/2})\tau_{x}^{1-3\alpha}\,.
\end{eqnarray*}
With the estimate $\frac{\sqrt{2}}{2}\leq \mathcal{E}_{H,\min}\leq C$,
we deduce
$$
\mathcal{E}_{\tau}(u_{1})= \mathcal{E}_{H,min}+
\mathcal{O}(\tau_{x}^{6\alpha}+ \tau_{x}^{1-3\alpha})\,.
$$
It's time to fix $\alpha$ to the value $\frac{1}{9}$ so that
$\tau_{x}^{6\alpha}=\tau_{x}^{1-3\alpha}
=\tau_{x}^{2/3}$
and
$$
\inf_{u\in H^{1}(\field{R}^{2})}\mathcal{E}_{\tau}(u)\leq
\mathcal{E}_{\tau}(u_{1})\leq \mathcal{E}_{H,min}+ \kappa \tau_{x}^{2/3}\,.
$$
\noindent\underline{Second step - Existence of a minimizer:} Once the function
$u_{1}\in H^{1}(\field{R}^{2})$ has been constructed as above, consider
$\tau_{x}<\tau_{0}$ with $\mathcal{E}_{H,min}+\kappa
\tau_{0}^{2/3}\leq \frac{1}{\ell_{V}^{2}\tau_{0}}$\,.
The functional
$$
\mathcal{E}_{\tau}(u)-\frac{1}{\ell_{V}^{2}\tau_{x}}\|u\|_{L^{2}}^{2}\,,
$$
is the sum of a convex strongly continuous functional (and therefore
weakly continuous) on
$H^{1}(\field{R}^{2})$ and a negative functional
$$
\langle u, \frac{1}{\ell_{V}^{2}\tau_{x}}\left[v(\tau_{x}^{1/2}.)-1\right]_{-}u\rangle\,.
$$
Due to the compact support of $v-1$, it is also continuous w.r.t the
weak topology on $H^{1}(\field{R}^{2})$\,. Out a minimizing sequence
$(u_{n})_{n\in \nz^{*}}$,
extract a weakly converging subsequence in $H^{1}(\field{R}^{2})$\,. The
weak limit, $u_{\infty}$, satisfies
$$
\mathcal{E}_{\tau}(u_{\infty})-
\frac{1}{\ell_{V}^{2}\tau_{x}}\|u_{\infty}\|_{L^{2}}^{2}=\lim_{k\to\infty}\mathcal{E}_{\tau}(u_{n_{k}})
-
\frac{1}{\ell_{V}^{2}\tau_{x}}\|u_{n_{k}}\|_{L^{2}}^{2}\,.
$$
with $\|u_{\infty}\|_{L^{2}}\leq 1$\,.
The same convergence holds also for the energy
$\mathcal{E}_{\tau}(u)-\frac{1}{2\ell_{V}^{2}\tau_{x}}\|u\|_{L^{2}}^{2}$, so
that actually $\|u_{\infty}\|_{L^{2}}=\lim_{k\to\infty}\|u_{n_{k}}\|_{L^{2}}=1$
and $u_{\infty}$ realizes the minimum of $\mathcal{E}_{\tau}(u)$ under
the constraint $\|u\|_{L^{2}}=1$\,.\\
The Euler-Lagrange equation can thus be written, with the stated
straightforward consequences.\\
\noindent\underline{Third step - a priori estimate for minimizers of
$\mathcal{E}_{\tau}$:}\\
Let $u\in H^{1}(\field{R}^{2})$ satisfy $\|u\|_{L^{2}}=1$ and
$\mathcal{E}_{\tau}(u)=\mathcal{E}_{\tau,min}\leq \mathcal{E}_{H}+\kappa \tau_{x}^{2/3}$\,.
Take two pairs $\tilde{\chi}=(\tilde{\chi}_{1},\tilde{\chi}_{2})$ and
$\chi'=(\chi'_{1},\chi'_{2})$, like
in our statement, and such that $\chi_{1}'\prec \tilde{\chi}_{1}$. The
identities \eqref{eq.IMSR} for $\mathcal{E}_{\tau}$ and \eqref{eq.IMSR} in
Lemma~\ref{le.compred} provide
\begin{eqnarray*}
&&\mathcal{E}_{\tau,min}\geq \mathcal{E}_{\tau}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)-C_{\tilde{\chi}}\tau_{x}^{2/9}
\\
&&
\mathcal{E}_{\tau,min}
\geq \mathcal{E}_{H}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)-C_{\tilde{\chi}}\tau_{x}^{2/9}
\\
&&\qquad-\frac{1}{4}\left[\mathcal{E}_{\tau}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)^{1/2}+
\mathcal{E}_{H}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)^{1/2}\right]\tau_{x}^{2/3}\,.
\end{eqnarray*}
The first line says
$$
\mathcal{E}_{\tau}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)
\leq \mathcal{E}_{\tau,min}+C_{\tilde{\chi}}\tau_{x}^{2/9}\leq
\mathcal{E}_{H,min}+ \kappa \tau_{x}^{2/3}+
C_{\tilde{\chi}}\tau_{x}^{2/9}\leq C_{\tilde{\chi}}'\,,
$$
which combined with the second line provides the uniform estimate
$$
\mathcal{E}_{\tau}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)\leq C''_{\tilde{\chi}}\,.
$$
Therefore $\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u$ is uniformly bounded in
$\mathcal{H}_{1}$ with respect to $\tau_{x}$\,.
Consider now the Euler-Lagrange equation
$$
\left[-\partial_{x}^{2}-(\partial_{y}-\frac{i}{2}\frac{x}{\sqrt{1+\tau_{x}x^{2}}})^{2}
+ \frac{v(\tau_{x}^{1/2}.)}{\ell_{V}^{2}\tau_{x}}+G|u|^{2}\right]u=\lambda_{u}u
$$
and write its local version for $u_{1}'=\chi'_{1}(\tau_{x}^{2/9}.)u$ in the
form
\begin{eqnarray}
\nonumber
&&
\left[-\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\right]u'_{1}
=\lambda_{u}u_{1}'-G|\tilde{u}_{1}|^{2}u'_{1}\\
&&
\label{eq.HHu1}\hspace{6cm}+f_{u}^{1}+f^{2}_{u}\,,
\\
\nonumber
\text{with}
&&f^{1}_{u}=-ix(\frac{1}{\sqrt{1+\tau_{x}x^{2}}}-1)\chi'(\tau_{x}^{1/9}.)\partial_{y}\tilde{u}_{1}
- \frac{x^{2}}{4}\frac{\tau_{x}x^{2}}{1+\tau_{x}x^{2}}u'_{1}\,,\\
\nonumber
\text{and}
&&f_{u}^{2}=
-2\tau_{x}^{1/9}(\nabla \chi_{1}')(\tau_{x}^{1/9}.).\nabla
\tilde{u}_{1}-\tau_{x}^{2/9}(\Delta\chi_{1}')(\tau_{x}^{1/9}.)\tilde{u}_{1}\,,
\end{eqnarray}
after setting
$\tilde{u}_{1}=\tilde{\chi}_{1}(\tau_{x}^{2/9}.)u$\,. Both functions,
$\tilde{u}_{1}$ and therefore $u'_{1}=\chi'_{1}(\tau_{x}^{2/9}.)\tilde{u}_{1}$
are uniformly estimated in $\mathcal{H}_{1}$ and therefore in
$H^{1}(\field{R}^{2})$\,.
From the embedding $H^{1}(\field{R}^{2})\subset L^{6}(\field{R}^{2})$, the term
$G|\tilde{u}_{1}|^{2}u'_{1}$ is uniformly bounded in $L^{2}(\field{R}^{2})$\,.
For the term $f_{u}^{1}$, the support condition $\supp
\chi'_{1}(\tau_{x}^{1/9}.)\subset \left\{|x|\leq\tau_{x}^{-1/9}\right\}$ imply
$$
\|f^{1}_{u}\|_{L^{2}}\leq C_{\chi'}(\tau_{x}^{1-1/3}+
\tau_{x}^{1-4/9})\leq C''_{\chi'}\,,
$$
while the estimate
$$
\|f_{u}^{2}\|\leq C_{\chi'}\tau_{x}^{1/9}\leq C_{\chi'}''
$$
is straightforward.
Hence the right-hand side of \eqref{eq.HHu1} is uniformly bounded in
$L^{2}(\field{R}^{2})$ and we have proved
\begin{equation}
\label{eq.estim1H2}
\|\chi'_{1}(\tau_{x}^{1/9}.)u\|_{\mathcal{H}_{2}}\leq C^{3}_{\chi'}
\end{equation}
for any good pair of cut-offs $\chi'=(\chi_{1}',\chi_{2}')$\,.
\\
\noindent\underline{Fourth step- accurate comparison of minimal energies:}\\
We already know $\mathcal{E}_{\tau,min}\leq
\mathcal{E}_{H,min}+\kappa\tau_{x}^{2/3}$ and we want to check the
reverse inequality. Consider a minimizer $u$ of $\mathcal{E}_{\tau}$ and
take two pairs of cut-off $\chi=(\chi_{1},\chi_{2})$ and
$\chi'=(\chi_{1}',\chi_{2}')$, such that $\chi_{1}\prec \chi_{1}'$\,.
The identity \eqref{eq.IMSH} for $\mathcal{E}_{\tau}$ and
\eqref{eq.IMSR} of Lemma~\ref{le.compred} used with $\chi$
imply
\begin{eqnarray}
\nonumber
&&
\mathcal{E}_{\tau,min}\geq
\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{1/9}.)u)+
\mathcal{E}_{\tau}(\chi_{2}(\tau_{x}^{1/9}.)u)
\\
\label{eq.ERmin1}
&&
\hspace{3cm}
-\tau_{x}^{2/9}\sum_{j=1}^{2}\int_{\field{R}^{2}}|(\nabla\chi_{j})(\tau_{x}^{1/9}.)|^{2}|u|^{2}\,,
\\
\nonumber
&&
\mathcal{E}_{\tau,min}
\geq \mathcal{E}_{H}(\chi_{1}(\tau_{x}^{1/9}.)u)
\\
&&\label{eq.ERmin2}
\hspace{3cm}-\tau_{x}^{2/9}\sum_{j=1}^{2}\int_{\field{R}^{2}}|(\nabla\chi_{j})(\tau_{x}^{1/9}.)|^{2}|u|^{2}+R(u)\,.
\end{eqnarray}
After setting
$u_{1}'=\chi_{1}'(\tau_{x}^{1/9}.)u$, we get the bound
$$
\int_{\field{R}^{2}}|(\nabla \chi_{j})(\tau_{x}^{1/9}.)|^{2}|u|^{2}
=
\int_{\field{R}^{2}}\frac{|(\nabla
\chi_{j})(\tau_{x}^{1/9}.)|^{2}}{|q|^{4}}||q|^{2}u'_{1}|^{2}\leq C_{\chi}\tau_{x}^{4/9}\,,
$$
while we already know from the third step the bounds
$\mathcal{E}_{\bullet}(\chi_{1}(\tau_{x}^{1/9}.)u)\leq C_{\chi}$ when
$\bullet$ stands for $\tau$ or $H$, which implies $|R(u)|\leq
C_{\chi}\tau_{x}^{2/3}$\,.
From \eqref{eq.ERmin1}, we deduce, as we did in the first step with
the energy $\mathcal{E}_{H}$,
$$
\|\chi_{2}(\tau_{x}^{1/9}.)u\|_{L^{2}}^{2}\leq
C_{\chi}\tau_{x}^{2/3}\quad
,\quad \|\chi_{1}(\tau_{x}^{1/9}.)u\|_{L^{2}}=1+\mathcal{O}(\tau_{x}^{2/3})\,,
$$
while the second line implies
$$
\mathcal{E}_{\tau,min}\geq
\left(1-C_{\chi}\tau_{x}^{2/3}\right)\mathcal{E}_{H,min}-C_{\chi}'\tau_{x}^{2/3}
\geq \mathcal{E}_{H,min}-C''_{\chi}\tau_{x}^{2/3}\,.
$$
\noindent\underline{Fifth step- accurate comparison of minimizers:}\\
The function
$u_{1}=\|\chi_{1}(\tau_{x}^{1/9}.)u\|_{L^{2}}^{-1}\chi_{1}(\tau_{x}^{1/9}.)u$,
satisfies
$$
\mathcal{E}_{H}(u_{1})\leq \mathcal{E}_{H,min}+ C_{\chi}\tau_{x}^{2/3}
$$
while $\chi_{1}(\tau_{x}^{1/9}.)u$ solves the equation \eqref{eq.HHu1} for some pair
$\tilde{\chi}=(\tilde{\chi}_{1},\tilde{\chi}_{2})$ such that
$\chi_{1}\prec \tilde{\chi}_{1}$ after replacing $(\chi', u_{1}')$
with $(\chi,\chi_{1}(\tau_{x}^{1/9}.)u)$\,. After normalization by
setting $\tilde{u}_{1}=\|\chi(\tau_{x}^{1/9}.)u\|_{L^{2}}^{-1}\chi_{1}(\tau_{x}^{1/9}.)u$ it becomes
\begin{eqnarray}
\label{eq.HHu2}
&&
H_{\ell_{V},G}u_{1}+G|u_{1}|^{2}u_{1}
=\lambda_{u}u_{1}+f_{u}^{1}+f^{2}_{u}+f^{3}_{u}\,,
\\
\nonumber
&&u_{1}=\|\chi(\tau_{x}^{1/9}.)u\|_{L^{2}}^{-1}\chi_{1}(\tau_{x}^{1/9}.)u\quad,\quad
\tilde{u}_{1}=\|\chi(\tau_{x}^{1/9}.)u\|_{L^{2}}^{-1}\tilde\chi_{1}(\tau_{x}^{1/9}.)u\,,\\
\nonumber
\text{with}
&&f^{1}_{u}=-ix(\frac{1}{\sqrt{1+\tau_{x}x^{2}}}-1)\chi(\tau_{x}^{1/9}.)
\partial_{y}\tilde{u}_{1}
- \frac{x^{2}}{4}\frac{\tau_{x}x^{2}}{1+\tau_{x}x^{2}}u_{1}\,,
\\
\nonumber
&&f_{u}^{2}=
-2\tau_{x}^{1/9}(\nabla \chi_{1})(\tau_{x}^{1/9}.).\nabla
\tilde{u}_{1}-\tau_{x}^{2/9}(\Delta\chi_{1})(\tau_{x}^{1/9}.)\tilde{u}_{1}\,,
\\
\nonumber
\text{and}
&&
\hspace{-0.5cm}f_{u}^{3}=G\|\chi_{1}(\tau_{x}^{1/9}.)u\|_{L^{2}}^{2}(|\tilde{u}_{1}|^{2}-|u_{1}|^{2})u_{1}+
G(1-\|\chi_{1}(\tau_{x}^{1/9}.)u\|_{L^{2}}^{2})|u_{1}|^{2}u_{1}\,.
\end{eqnarray}
The estimate $\eqref{eq.estim1H2}$,
for any new good pair
$\chi'=(\chi_{1}',\chi_{2}')$ such that
$\chi_{1}\prec\tilde{\chi}_{1}\prec\chi'_{1}$,
implies that the terms $f^{1}_{u}$ and $f_{u}^{2}$ of the right-hand side of \eqref{eq.HHu2}
have an
$L^{2}$-norm
of order $\tau_{x}^{2/3}$\,. For the third term the estimate
\eqref{eq.estim1H2} also implies that
$$
(|\tilde{u}_{1}|^{2}-|u_{1}|^{2})u_{1}=\left(1-\chi_{1}^{2}(\tau_{x}^{1/9}.)\right)|\tilde{u_{1}}|^{2}u_{1}\,,
$$
has an $L^{2}$-norm of order $\mathcal{O}(\tau_{x}^{2/3})$ (use the
$L^{\infty}$ bound for $|\tilde{u}_{1}|^{2}$ with
$\||q|^{2}u_{1}\|_{L^{2}}\leq C_{\chi}$).
We conclude by applying Proposition~\ref{pr.harmH2}\,.
{\hfill $\qed${}
We end this section with a comparison property similar to
Proposition~\ref{pr.harmH2}.
\begin{proposition}
\label{pr.appH2}
Let $\ell_{V},G$ be fixed positive numbers and take $u\in
H^{1}(\field{R}^{2})$ such that
$$
\mathcal{E}_{\tau}(u)\leq \mathcal{E}_{\tau,min}+C_{\ell_{V,G}}\tau_{x}^{2/3}
$$
and which solves
$$
-\partial_{x}^{2}u-(\partial_{y}-i\frac{x}{\sqrt{1+\tau_{x}x^{2}}})^{2}+
\frac{v(\tau_{x}^{1/2}.)}{\ell_{V}^{2}\tau_{x}}u+G|u|^{2}u=\lambda_{u}u+ r_{u}
$$
with $\lambda_{u}\in \field{R}$ and $\|r_{u}\|_{L^{2}}\leq 1$\,.
Then for any pair
$\chi=(\chi_{1},\chi_{2})\in \mathcal{C}^{\infty}_{b}(\field{R}^{2})$ so
that $\chi_{1}^{2}+\chi_{2}^{2}=1$ with $\supp
\chi_{1}\subset\left\{x^{2}+y^{2}<1\right\}$ and $\chi_{1}\equiv 1$
in $\left\{x^{2}+y^{2}\leq 1/2\right\}$\,, there exists
$\tau_{\chi,\ell_{V},G}$ and $C_{\chi,\ell_{V}}$ such that
\begin{eqnarray}
\label{eq.estimu2a}
&& \|\chi_{2}(\tau_{x}^{1/9}.)u\|^{2}_{L^{2}}\leq
C_{\chi,\ell_{V},G}\tau_{x}^{2/3}\\
&&
\label{eq.estimEnu1a}
|\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{1/9}.)u)-\mathcal{E}_{H,min}|\leq C_{\chi,\ell_{V},G}\tau_{x}^{2/3}
\\
\label{eq.estimH2u1a}
&& d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{1/9}.)u,
\mathrm{Argmin}~\mathcal{E}_{H})\leq
C_{\chi,\ell_{V},G}(\tau_{x}^{2\nu_{\ell_{V,G}}/3}+\|r_{u}\|_{L^{2}})\,,
\end{eqnarray}
when $\tau_{x}< \tau_{\chi,\ell_{V},G}$ and
where $\nu_{\ell_{V},G}\in (0,\frac{1}{2}]$ is the exponent given in Proposition~\ref{pr.harmapp}.
\end{proposition}
\noindent{\bf Proof: \ }{}
The analysis follows essentially the same line as the study of the
minimizers of $\mathcal{E}_{\tau}$ in the proof of
Proposition~\ref{pr.redHR}.
By taking two pairs $\chi'=(\chi'_{1},\chi'_{2})$ and
$\tilde{\chi}=\tilde(\tilde{\chi}_{1},\tilde{\chi}_{2})$ such that
$\chi'_{1}\prec \tilde{\chi_{1}}$, we obtain successively like in the Third~Step
in the proof of Proposition~\ref{pr.redHR}~:
\begin{itemize}
\item $\mathcal{E}_{\tau}(\tilde{\chi}_{1}(\tau_{x}^{1/9}.)u)+
\mathcal{E}_{\tau}(\tilde{\chi}_{2}(\tau_{x}^{1/9}.)u) \leq
C_{\tilde\chi}$\,;
\item $\left[-\partial_{x}^{2}-(\partial_{y}-\frac{ix}{2})^{2}
+ \frac{x^{2}+y^{2}}{\ell_{V}^{2}}\right]u'_{1}
=\lambda_{u}u_{1}'-G|\tilde{u}_{1}|^{2}u'_{1}+f_{u}^{1}+f^{2}_{u}+r_{u}$
where $u_{1}'$,$\tilde{u}_{1}$, $f_{u}^{1,2}$ have the same
expressions as in \eqref{eq.HHu1};
\item $\|\chi'_{1}(\tau_{x}^{1/9}.)u\|_{\mathcal{H}_{2}}\leq C_{\chi'}$
owing to $\|r_{u}\|_{L^{2}}\leq 1$\,.
\end{itemize}
From the last estimate, the refined comparison of energies like in
the Fourth~Step gives for a pair $\chi=(\chi_{1},\chi_{2})$ such that
$\chi_{1}\prec \chi'_{1}$:
\begin{eqnarray*}
\|\chi_{2}(\tau_{x}^{1/9}.)u\|_{L^{2}}^{2}\leq C_{\chi}\tau_{x}^{2/3}\quad
\mathcal{E}_{H}(u_{1})\leq \mathcal{E}_{H,min}+C_{\chi}\tau_{x}^{2/3}\,,
\end{eqnarray*}
with
$u_{1}=\|\chi_{1}(\tau_{x}^{1/9}.)u\|^{-1}\chi_{1}(\tau_{x}^{1/9}.)u$
and $C_{\chi}\tau_{x}^{2/3}\leq 1$ for $\tau_{x}\leq \tau_{\chi}$\,.
The equation~\eqref{eq.HHu2} is replaced by
$$
H_{\ell_{V},G}u_{1}+G|u_{1}|^{2}u_{1}=\lambda_{u}u_{1}+f_{u}^{1}+f_{u}^{2}+f_{u}^{3}+\|\chi_{1}(\tau_{x}^{1/9}.)u\|^{-1}r_{u}
$$
without changing the expressions of $f_{u}^{1,2,3}$\,. Again the
estimate $\|\chi'_{1}(\tau_{x}^{1/9}.)u\|_{\mathcal{H}_{2}}\leq C_{\chi'}$ is used with various cut-offs
$\chi_{1}'$, in order to get
$\|f^{1}_{u}+f^{2}_{u}+f^{3}_{u}\|_{L^{2}}=\mathcal{O}(\tau_{x}^{2/3})$\,. We
conclude with the help of Proposition~\ref{pr.harmH2} applied to
$u_{1}$\,.
{\hfill $\qed${}
\section{Analysis of the complete minimization problem}
\label{se.compmini}
We consider the complete minimization problem for the energy
$$
\mathcal{E}_{\varepsilon}(\psi)=\langle
\psi,H_{Lin}\psi\rangle+\frac{G_{\varepsilon,\tau}}{2}\int|\psi^{4}|
= \varepsilon^{2+2\delta}\tau_{x}\left[\langle
\psi,\varepsilon^{-2-2\delta}\tau_{x}^{-1}
H_{Lin}\psi\rangle+\frac{G}{2}\int|\psi^{4}|\right]
$$
and compare its solutions to the minimization of the reduced energies
$\mathcal{E}_{\tau}$ and
$\mathcal{E}_{H}$, introduced in the previous sections. We work with
$\tau_{y}=1$, $\tau_{x}\to 0$, $\varepsilon\to 0$, while $\ell_{V},G$
and $\delta\in (0,\delta_{0}]$ are fixed.
The analysis follows the same lines as the proof of Proposition~\ref{pr.redHR}.
\subsection{Upper bound for
$\inf\left\{\mathcal{E}_{\varepsilon}(\psi), \|\psi\|_{L^{2}}=1\right\}$}
\label{se.uppcomplete}
The potential $V_{\varepsilon}$ is chosen according
to \eqref{eq.defVepsIntro}-\eqref{eq.defvIntro} while $\ell_{V}$ and $G$ are
fixed. The parameter $\tau_{x}$ is assumed to be smaller than
$\tau_{\ell_{V},G}$ so that the minimal energy
$\mathcal{E}_{\tau,min}(\tau_{x})$
of $\mathcal{E}_{\tau}$ is achieved (see Proposition~\ref{pr.redHR}) and
$|\mathcal{E}_{\tau,min}(\tau_{x})-\mathcal{E}_{H,min}|\leq
C_{\ell_{V},G}\tau_{x}^{2/3}$\,.
Moreover Proposition~\ref{pr.redHR} also says that by truncating an
element of $\mathrm{Argmin}~\mathcal{E}_{\tau}$, one can find
$a_{-}\in \mathcal{H}_{2}$ such that
\begin{equation}
\label{eq.choixtest}
\|a_{-}\|_{L^{2}}=1\,,\quad |\mathcal{E}_{\tau}(a_{-})-\mathcal{E}_{H,min}|\leq
C_{\ell_{V},G}\tau_{x}^{2/3}
\quad\text{and}\quad
\|a_{-}\|_{\mathcal{H}_{2}}\leq C_{\ell_{V},G}\,.
\end{equation}
\begin{proposition}
\label{pr.bornesup}
Under the above assumptions, take $\psi=
\hat{U}\begin{pmatrix}
0\\e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$ where $a_{-}$ satisfies \eqref{eq.choixtest} and
$\hat{U}=U(q,\varepsilon D_{q},\tau,\varepsilon)$ is the unitary
operator introduced in Theorem~\ref{th.BornOp}.
The estimate
\begin{equation}
\label{bornesup}
\left|\mathcal{E}_{\varepsilon}(\psi)- \varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau}(a_{-})
\right|\leq C_{\ell_{V},G}\varepsilon^{2+4\delta}\,.
\end{equation}
hold uniformly w.r.t $\tau_{x}\in (0,\tau_{\ell_{V},G}]$ and
$\delta\in (0,\delta_{0}]$\,.
\end{proposition}
\noindent{\bf Proof: \ }
Let us compare first the linear part by estimating
\begin{multline*}
\left|\langle \psi\,,
H_{Lin}\psi\rangle-\varepsilon^{2+2\delta}\tau_{x}\langle a_{-}\,,
\left[-\partial_{x}^{2}-(\partial_{y}-i
\frac{x}{2\sqrt{1+\tau_{x}x^{2}}})^{2}+\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}\tau_{x}}\right]a_{-}\rangle\right|
\\
=\left|
\langle \psi\,,
(H_{Lin}-\varepsilon^{2+2\delta}\tau_{x}U^{*}H_{BO}U )\psi\rangle
\right|
\end{multline*}
By Proposition~\ref{pr.adiabexpl2}, it suffices to estimate
$$
\|(1-\hat{\chi})\psi\|_{L^{2}}\quad\text{and}\quad
\|\sqrt{\tau_{x}}|\varepsilon D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}
$$
for some given cut-off function $\chi\in
\mathcal{C}^{\infty}_{0}(\field{R})$ with
$\hat{\chi}=\chi(\tau_{x}\varepsilon^{2}D_{q}^{2})$\,.
Notice
$$
e^{i\frac{y}{2\sqrt{\tau_{x}}}}(\sqrt{\tau_{x}}\varepsilon D_{q})e^{-i\frac{y}{2\sqrt{\tau_{x}}}}
=
\begin{pmatrix}
\sqrt{\tau_{x}}\varepsilon D_{x}\\
\sqrt{\tau_{x}}\varepsilon D_{y}-\frac{\varepsilon}{2}
\end{pmatrix}
$$
Hence by using the functional calculus of $\sqrt{\varepsilon}D_{q}$,
we can say that there exists a cut-off $\chi'\in
\mathcal{C}^{\infty}_{0}(-r_{\gamma}^{2},r_{\gamma}^{2})$, with
$\chi'\equiv 1$ around $0$ and $\chi'\prec\chi$ such that
\begin{eqnarray*}
&&
e^{i\frac{y}{2\sqrt{\tau_{x}}}}(1-\hat{\chi})^{2}e^{-i\frac{y}{2\sqrt{\tau_{x}}}}\leq
(1-\hat{\chi'})^{2}\,,
\\
\text{and}
&&
e^{i\frac{y}{2\sqrt{\tau_{x}}}}(\tau_{x}\varepsilon^{2}|D_{q}|^{2})(1-\hat{\chi})^{2}e^{-i\frac{y}{2\sqrt{\tau_{x}}}}
\leq 2(\tau_{x}\varepsilon^{2}|D_{q}|^{2})(1-\hat{\chi'})^{2}+ 2(1-\hat{\chi'})^{2}
\end{eqnarray*}
as soon as $\varepsilon\leq \varepsilon_{0}\leq 1$, for a convenient choice
of $\varepsilon_{0}$ and $r_{\gamma}$\,.
By using $\|a\|_{H^{2}(\field{R}^{2})}\leq C_{\ell_{V},G}$, we deduce
$$
\|(1-\hat{\chi})\psi\|_{L^{2}}^{2}\leq
\|(1-\hat{\chi'})a_{-}\|_{L^{2}}^{2}\leq
C_{\ell_{V},G}\tau_{x}^{2}\varepsilon^{4}\,,
$$
and
\begin{eqnarray*}
\|(\sqrt{\tau_{x}}\varepsilon|D_{q}|)(1-\hat{\chi})\psi\|_{L^{2}}^{2}
&\leq &
2\|(\sqrt{\tau_{x}}\varepsilon|D_{q}|)(1-\hat{\chi'})a_{-}\|_{L^{2}}^{2}
+2\|(1-\hat{\chi'})a_{-}\|_{L^{2}}^{2}\\
&\leq& C_{\ell_{V},G}\tau_{x}\varepsilon^{2}\,.
\end{eqnarray*}
By Proposition~\ref{pr.adiabexpl2}, we obtain
\begin{multline}
\label{eq.erreurenerlin}
\left|
\langle \psi\,,
(H_{Lin}-\varepsilon^{2+2\delta}\tau_{x}U^{*}H_{BO}U )\psi\rangle
\right|\\
\leq C_{\ell_{V},G}\left[\varepsilon^{2+4\delta}+
\varepsilon^{5+2\delta}\tau_{x}^{2}+
\varepsilon^{4+2\delta}\tau_{x}^{3/2}\right]
\leq C_{\ell_{V},G}'\varepsilon^{2+4\delta}\,.
\end{multline}
For the nonlinear part of the energy,
Proposition~\ref{pr.L4err} gives
$$
(1-C\varepsilon^{1+2\delta})\int_{\field{R}^{2}}|a_{-}|^{4}\leq
\int_{\field{R}^{2}}|\psi|^{4}\leq (1+C\varepsilon^{1+2\delta})\int_{\field{R}^{2}}|a_{-}|^{4}
$$
and the bound,
$\int_{\field{R}^{2}}|e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}|^{4}=\int_{\field{R}^{2}}|a_{-}|^{4}\leq
C_{\ell_{V},G}$, leads to
$$
\left|\frac{G\tau_{x}\varepsilon^{2+2\delta}}{2}
\int_{\field{R}^{2}}|\psi|^{4}-\varepsilon^{2+2\delta}\frac{G\tau_{x}}{2}\int_{\field{R}^{2}}|a_{-}|^{4}\right|
\leq
C_{\ell_{V},G}\varepsilon^{2+2\delta}\tau_{x}\times \varepsilon^{1+2\delta}\,,
$$
which is smaller than the error term for the linear part.
This ends the proof of \eqref{bornesup}\,.
{\hfill $\qed$
\begin{remarque}
\label{re.epstau}
\begin{itemize}
\item The energy
$\mathcal{E}_{\tau}(a_{-})=\mathcal{E}_{H,min}+\mathcal{O}(\tau_{x}^{2/3})$\,. Therefore
the error given by \eqref{bornesup} is relevant, as compared with
the energy scale of
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau,min}$, when
\begin{equation}
\label{eq.compepstaurel1}
\varepsilon^{2\delta}\leq c_{\ell_{V},G}\tau_{x}\,,
\end{equation}
with $c_{\ell_{V},G}$ small enough,
and accurate when
\begin{equation}
\label{eq.compepstaurel2}
\varepsilon^{2\delta}\leq C_{\ell_{V},G}\tau_{x}^{5/3}\,.
\end{equation}
Remember that the constants $c_{\ell_{V},G}$ and
$C_{\ell_{V},G}$ depend also on $\delta_{0}$, when
$\delta\in (0,\delta_{0}]$\,.
\item It is interesting to notice that the worst term in the right-hand
side of \eqref{eq.erreurenerlin} comes from the error of order
$\mathcal{O}(\varepsilon^{2+2\delta})$ in the Born-Oppenheimer
approximation. There seems to be no way to get an additional factor
$\tau_{x}^{\alpha}$ with $\alpha>0$ because the initial problem is
rapidly oscillatory in the $y$-variable in a $\tau_{x}$-dependent
scale. This can be seen on the gain associated with
the metric $g_{\tau}$, for $\tau_{y}=1$ and $\tau_{x}>0$, which is simply
$\langle \sqrt{\tau}_{x}p\rangle$ or essentially $1$ when $p$ is small.
\end{itemize}
\end{remarque}
\subsection{Existence of a minimizer for $\mathcal{E}_{\varepsilon}$}
\label{se.exisEeps}
With the choice \eqref{eq.defVepsIntro}-\eqref{eq.defvIntro} of the potential
$V_{\varepsilon,\tau}$, the linear Hamiltonian $H_{Lin}$ can be
written
\begin{eqnarray*}
&&
H_{Lin}=-\varepsilon^{2+2\delta}\tau_{x}\Delta + u_{0}(q,\tau)
\begin{pmatrix}
2\sqrt{1+\tau_{x}x^{2}}&0\\
0&0
\end{pmatrix}u_{0}(q,\tau)^{*}\\
&&\hspace{5cm}+\varepsilon^{2+2\delta}\frac{v(\tau_{x}^{1/2}.)}{\ell_{V}^{2}}
-\varepsilon^{2+2\delta}W_{\tau}(x,y)\,,\\
\text{with}
&&W_{\tau}(x,y)=\frac{\tau_{x}^{2}}{(1+\tau_{x}x^{2})^{2}}+
\frac{1}{1+\tau_{x}x^{2}}\,.
\end{eqnarray*}
For $t\leq \frac{\varepsilon^{2+2\delta}}{2\ell_{V}^{2}}$, the
negative part
$\left(\varepsilon^{2+2\delta}\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}-t\right)_{-}$ is compactly supported. Set
\begin{multline*}
H_{Lin,t,+}=-\varepsilon^{2+2\delta}\tau_{x}\Delta + u_{0}(q,\tau)
\begin{pmatrix}
2\sqrt{1+\tau_{x}x^{2}}&0\\
0&0
\end{pmatrix}u_{0}(q,\tau)^{*}
\\
+
\left(\varepsilon^{2+2\delta}\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}-t\right)_{+}\,,
\end{multline*}
so that
\begin{multline}
\label{eq.decompEeps}
\mathcal{E}_{\varepsilon}(\psi)-t\|\psi\|^{2}=\langle
\psi\,,H_{Lin,t,+}\psi\rangle
+\frac{G_{\varepsilon,\tau}}{2}\int|\psi|^{4}
\\
+ \langle
\psi\,,\,\varepsilon^{2+2\delta}\left[\left(\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}-\frac{t}{\varepsilon^{2+2\delta}}\right)_{-}
-W_{\tau}\right]\psi\rangle\,.
\end{multline}
\begin{proposition}
\label{pr.existEeps}
Assume $\varepsilon\leq\varepsilon_{\ell_{V},G}$ and $\tau_{x}\leq
\tau_{\ell_{V},G}$ with $\tau_{\ell_{V},G}$ small enough. Then the infimum
$$
\inf\{\mathcal{E}_{\varepsilon}(u), \quad u\in H^{1}(\field{R}^{2}), \|u\|_{L^{2}}=1\}
$$
is achieved. Any element $\psi$ of
$\mathrm{Argmin}~\mathcal{E}_{\varepsilon}$ solves an Euler-Lagrange
equation
$$
H_{Lin}\psi+G_{\varepsilon,\tau}|\psi|^{2}\psi=\lambda_{\psi}\psi
$$
with the estimates
\begin{eqnarray*}
&&
|\mathcal{E}_{\varepsilon}(\psi)|+|\lambda_{\psi}|\leq C_{\ell_{V},G}\varepsilon^{2+2\delta}
\quad,\quad
\|(1+\tau_{x}|D_{q}|^{2})^{1/2}\psi\|_{L^{2}}\leq C_{\ell_{V},G}\,,
\\
&&
\begin{array}[c]{ll}
\mathcal{E}_{\varepsilon}(\psi)=\mathcal{E}_{\varepsilon,min}&\leq
\varepsilon^{2+2\delta}\left[\mathcal{E}_{\tau,min}+C_{\ell_{V},G}\varepsilon^{2\delta})\right]
\\
&\leq
\varepsilon^{2+2\delta}\left[\tau_{x}\mathcal{E}_{H,min}+
C_{\ell_{V},G}'(\tau_{x}^{5/3}+\varepsilon^{2\delta})\right]\,.
\end{array}
\end{eqnarray*}
\end{proposition}
\noindent{\bf Proof: \ }{}
From Proposition~\ref{pr.bornesup}, we know that
$$
\inf_{\|u\|_{L^{2}}=1}\mathcal{E}_{\varepsilon}(u)\leq
\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau,min}+
C_{\ell_{V},G}\varepsilon^{2+4\delta}=:\frac{t}{2}\,.
$$
For $\varepsilon\leq \varepsilon_{\ell_{V},G}$ and $\tau_{x}\leq
\tau_{\ell_{V},G}$, $t$ smaller than
$\frac{\varepsilon^{2+2\delta}}{2\ell_{V}^{2}}$\,. Consider the
decomposition \eqref{eq.decompEeps} for the energy
$\mathcal{E}_{\varepsilon}(u)-t\|u\|_{L^{2}}^{2}$\,.
By the same argument (convexity of the positive part and compactness of
the negative part) as we used for $\mathcal{E}_{\tau}$ in the proof of
Proposition~\ref{pr.redHR} (second step), a weak limit of an extracted
sequence of minimizers in $H^{1}(\field{R}^{2})$ is a minimum for
$\mathcal{E}_{\varepsilon}$ on $\left\{\|u\|_{L^{2}}=1\right\}$\,.\\
A element $\psi$ of $\mathrm{Argmin}~\mathcal{E}_{\varepsilon}$ satisfies
\begin{eqnarray*}
\varepsilon^{2+2\delta}\tau_{x}\left[\langle \psi\,,\, -\Delta
\psi\rangle+\frac{G}{2}\int|\psi|^{4}\right]
&&\leq \langle \psi\,,\, H_{Lin,t, +}\psi\rangle
+\frac{G_{\varepsilon,\tau}}{2}\int|\psi|^{4}
\\
&&\hspace{-2cm}\leq \mathcal{E}_{\varepsilon,min}-\langle
\psi\,,\,\varepsilon^{2+2\delta}\left[\left(\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}-\frac{t}{\varepsilon^{2+2\delta}}\right)_{-}
-W_{\tau}\right]\psi\rangle
\\
&&\leq
\frac{t}{2}+t+
(1+\tau_{x}^{2})\varepsilon^{2+2\delta}
\leq C_{\ell_{V},G}'\varepsilon^{2+2\delta}\,,
\end{eqnarray*}
by recalling $v\geq 0$ for the last line.
This implies
$$
\|\psi\|_{H^{1}}^{2}\leq
C_{\ell_{V},G}'\tau_{x}^{-1}\quad,\quad
\|\psi\|_{L^{4}}^{4}\leq \frac{2C_{\ell_{V},G}'}{G}\tau_{x}^{-1}\,,
$$
and by interpolation with $\|u\|_{L^{6}}\leq C\|\nabla
u\|_{L^{2}}^{2/3}\|u\|_{L^{2}}^{1/3}$,
$\|\psi\|_{L^{6}}=\mathcal{O}(\tau_{x}^{-1/3})$\,. The first
inequality with $\|\psi\|_{L^{2}}=1$, gives
$\|(1+\tau_{x}|D_{q}|^{2})^{1/2}\psi\|_{L^{2}}=\mathcal{O}(1)$\,.\\
The Euler-Lagrange equation
$$
H_{Lin}\psi+G_{\varepsilon,\tau}|\psi|^{2}\psi=\lambda\psi
$$
implies
\begin{eqnarray*}
|\lambda|&\leq &
2\left[\langle \psi\,,\, H_{Lin,t, +}\psi\rangle
+\frac{G_{\varepsilon,\tau}}{2}\int|\psi|^{4}\right]
\\
&&-\langle
\psi\,,\,\varepsilon^{2+2\delta}\left[\left(\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}-\frac{t}{\varepsilon^{2+2\delta}}\right)_{-}
-W_{\tau}\right]\psi\rangle
\leq C''_{\ell_{V},G}\varepsilon^{2+2\delta}\,.
\end{eqnarray*}
Similarly, the lower bound $\mathcal{E}_{\varepsilon}(\psi)\geq
-2\varepsilon^{2+2\delta}$ is due to $W_{\tau}\geq -2$\,.
The upper bound of
$\mathcal{E}_{\varepsilon}(\psi)=\mathcal{E}_{\varepsilon,min}$ comes
from Proposition~\ref{pr.bornesup} and \eqref{eq.choixtest}.
{\hfill $\qed${}
\subsection{Comparison of minimal energies between
$\mathcal{E}_{\varepsilon}$ and $\mathcal{E}_{\tau}$}
\label{se.accminen}
In this subsection, we specify a priori estimates for the minimizers of
$\mathcal{E}_{\varepsilon}$ and compare the energies
$\mathcal{E}_{\varepsilon,min}$ and $\mathcal{E}_{\tau,min}$ without
imposing relations between $\varepsilon^{2\delta}$ and
$\tau_{x}$\,. This is not necessary at this level, if one uses
carefully bootstrap arguments.
\begin{proposition}
\label{pr.compen}
Let $V_{\varepsilon,\tau}$ be given by
\eqref{eq.defVepsIntro}-\eqref{eq.defvIntro} and assume $\tau_{x}\leq
\tau_{\ell_{V},G}$ and $\varepsilon\leq \varepsilon_{\ell_{V},G}$ so
that $\mathcal{E}_{\varepsilon}$ admits a ground state according to
Proposition~\ref{pr.existEeps}\,.
The operator $\hat{U}$ is the unitary
transform provided by Theorem~\ref{th.BornOp} and an element
$\psi\in \mathrm{Argmin}~\mathcal{E}_{\varepsilon}$ is written
$\hat{U}
\begin{pmatrix}
e^{+i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$\,, with $a=
\begin{pmatrix}
a_{+}\\alpha_{-}
\end{pmatrix}
$\,.
Then, the estimates
\begin{eqnarray}
\label{eq.H1a}
&&\|(1+\tau_{x}|D_{q}|^{2})^{1/2}a\|_{L^{2}}\leq C_{\ell_{V},G}\,,\\
\label{eq.L2a+}
&&
\|a_{+}\|_{L^{2}}^{2}+G_{\varepsilon,\tau}\int|a|^{4}\leq
C_{\ell_{V},G}\varepsilon^{2+4\delta}+\mathcal{E}_{\varepsilon,min}\,,\\
\label{eq.ena-}
&&
|\mathcal{E}_{\varepsilon}(\psi)-\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau}(a_{-})|\leq
C_{\ell_{V},G}\varepsilon^{2+4\delta}\,,\\
&&
\label{eq.compEmin}
|\mathcal{E}_{\varepsilon,min}-\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau,min}|\leq C_{\ell_{V},G}\varepsilon^{2+4\delta}\,,
\end{eqnarray}
hold with right-hand
sides which can be
replaced
by $C_{\ell_{V},G}\varepsilon^{2+2\delta}\tau_{x}$ when
$\varepsilon^{2\delta}\leq c_{\ell_{V},G}\tau_{x}$\,.
\end{proposition}
\noindent{\bf Proof: \ }{}
For $\tilde{a}=
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}
$ and $a=\begin{pmatrix}
a_{+}\\
a_{-}
\end{pmatrix}$, the norms $\|(1+\tau_{x}|D_{q}|^{2})^{1/2}a\|_{L^{2}}$ and
$\|(1+\tau_{x}|D_{q}|^{2})^{1/2}\tilde{a}\|_{L^{2}}$ are uniformly equivalent
because
$$
e^{\pm
i\frac{y}{2\sqrt{\tau_{x}}}}(\sqrt{\tau_{x}}D_{y})e^{\mp
i\frac{y}{2\sqrt{\tau_{x}}}}
=(\sqrt{\tau_{x}}D_{y})\mp\frac{1}{2}\,.
$$
Hence it suffices to
estimate $\sqrt{\tau_{x}}\varepsilon D_{q}\hat{U}^{*}\psi$\,. Remember that
$\hat{U}=U(q,\varepsilon D_{q},\tau,\varepsilon)$ with $U\in
S_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, while $\sqrt{\tau_{x}}p\in
S_{u}(\langle\sqrt{\tau_{x}}p\rangle,g_{\tau};\field{R}^{2})$\,.
With $\|(1+\tau_{x}
|D_{q}|^{2})^{1/2}\psi\|_{L^{2}}\leq C_{\ell_{V},G}$ in
Proposition~\ref{pr.existEeps}, simply compute:
\begin{equation*}
\sqrt{\tau_{x}}\varepsilon D_{q}\hat{U}^{*}\psi
=\varepsilon \hat{U}^{*}\sqrt{\tau_{x}} D_{q}\psi +
\left[\sqrt{\tau_{x}}\varepsilon D_{q}, \hat{U}^{*}\right]\psi
=\mathcal{O}(\varepsilon)\quad\text{in}~L^{2}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2})\,,
\end{equation*}
and divide by $\varepsilon$ for \eqref{eq.H1a}.\\
In order to compare the energies, consider first the linear part by writing
\begin{multline*}
\left|\langle \psi\,,
H_{Lin}\psi\rangle-\varepsilon^{2+2\delta}\tau_{x}\left[\langle
a_{+}\,,\,\hat{H}_{+}a_{+}\rangle+ \langle a_{-}\,,\,\hat{H}_{-}a_{-}\rangle\right]\right|
=
\\
\left|
\langle \tilde{a}\,,
(\hat{U}^{*}H_{Lin}\hat{U}-\varepsilon^{2+2\delta}H_{BO})\tilde{a}\rangle
\right|\,,
\end{multline*}
with
\begin{eqnarray*}
\tilde{a}&=&
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}\,,
\\
H_{BO}&=&\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}\hat{H}_{+}e^{-i\frac{y}{2\sqrt{\tau_{x}}}}&0\\
0&e^{-i\frac{y}{2\sqrt{\tau_{x}}}}\hat{H}_{-}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}
\end{pmatrix}\,,
\\
\hat{H}_{\pm}&=&-\partial_{x}^{2}-(\partial_{y}\pm i
\frac{x}{2\sqrt{1+\tau_{x}x^{2}}})^{2}+\frac{v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}\tau_{x}}+\frac{(1\pm
1)}{\varepsilon^{2+2\delta}\tau_{x}} (1+\tau_{x}x^{2})^{1/2}\,.
\end{eqnarray*}
Here it is convenient to write
\begin{eqnarray}
\nonumber
&&\varepsilon^{2+2\delta}\tau_{x}H_{BO}= \varepsilon^{2\delta}
\hat{B}
+
\begin{pmatrix}
2(1+\tau_{x}x^{2})^{1/2}\\
0
\end{pmatrix}\,,
\\
\label{eq.defm1}
&&
\hat{B}= \begin{pmatrix}
\hat{B}_{+}& 0\\
0&\hat{B}_{-}
\end{pmatrix}\quad,\quad
\hat{B}_{\pm}=
B_{\pm}(q,\varepsilon D_{q},\tau,\varepsilon)\,,\\
\label{eq.defm2}
&&
\hspace{-2cm}
B_{\pm}(q,p,\tau,\varepsilon)=\tau_{x}p_{x}^{2}+
(\sqrt{\tau_{x}}p_{y}\pm
\frac{\varepsilon}{2}(\frac{\sqrt{\tau_{x}}x}{\sqrt{1+\tau_{x}x^{2}}}
-\sqrt{\tau_{x}}))^{2}
+\frac{\varepsilon^{2}v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}\,.
\end{eqnarray}
With the notation
$\hat{\chi}=\chi(\tau_{x}\varepsilon^{2}|D_{q}|^{2})$ of
Proposition~\ref{pr.adiabexpl2}, Lemma~\ref{le.ellipt} below says in
particular
\begin{eqnarray*}
&&
\|(1-\hat{\chi})\psi\|_{L^{2}}^{2}\leq C_{\ell_{V},G}\left[\langle
\tilde{a}, \hat{B}\tilde{a}\rangle + \varepsilon^{1+2\delta}\|\tilde{a}\|_{L^{2}}^{2}\right]\,,
\\
&&
\|\sqrt{\tau_{x}}\varepsilon|D_{q}|(1-\hat{\chi})\psi\|_{L^{2}}^{2}
\leq C_{\ell_{V},G}\left[\langle
\tilde{a}, \hat{B}\tilde{a}\rangle +
\varepsilon^{1+2\delta}\|\tilde{a}\|_{L^{2}}^{2}\right]\,.
\end{eqnarray*}
Proposition~\ref{pr.adiabexpl2} yields
$$
\left|
\langle \tilde{a}\,,
\left(\hat{U}^{*}H_{Lin}\hat{U}-\varepsilon^{2\delta}\hat{B}-
\begin{pmatrix}
2(1+\tau_{x}x^{2})^{1/2}\\
0
\end{pmatrix}\right)\tilde{a}\rangle
\right|\leq
C_{\ell_{V},G}\left[\varepsilon^{2+4\delta}+\varepsilon^{1+2\delta}\langle\tilde{a}\,,\,
\hat{B}\tilde{a}\rangle\right]\,.
$$
For the nonlinear part of the energy,
Proposition~\ref{pr.L4err} gives
$$
(1-C\varepsilon^{1+2\delta})\int_{\field{R}^{2}}|\psi|^{4}\leq
\int_{\field{R}^{2}}|\tilde{a}|^{4}=\int_{\field{R}^{2}}|a|^{4}\leq (1+C\varepsilon^{1+2\delta})\int_{\field{R}^{2}}|\psi|^{4}
$$
with $\tilde{a}=
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}
$ and $a=\begin{pmatrix}
a_{+}\\
a_{-}
\end{pmatrix}$\,.
The bound
$\int_{\field{R}^{2}}|\psi|^{4}\leq
C_{\ell_{V},G}\tau_{x}^{-1}$ coming from
$\mathcal{E}_{\varepsilon}(\psi)=\mathcal{O}(\varepsilon^{2+2\delta})$
with $\varepsilon^{2+2\delta}W_{\tau}=\mathcal{O}(\varepsilon^{2+2\delta})$, leads to
$$
\left|\frac{G\tau_{x}\varepsilon^{2+2\delta}}{2}
\int_{\field{R}^{2}}|\psi|^{4}-\frac{G\tau_{x}\varepsilon^{2+2\delta}}{2}\int_{\field{R}^{2}}|\tilde{a}|^{4}\right|
\leq
C_{\ell_{V},G}\varepsilon^{2+2\delta}\tau_{x}\times
\varepsilon^{1+2\delta}\tau_{x}^{-1}
=C_{\ell_{V},G}\varepsilon^{3+4\delta}\,,
$$
which is smaller than the error term for the linear part.
We have proved
\begin{multline*}
\left|\mathcal{E}_{\varepsilon}(\psi)-\varepsilon^{2\delta}
\langle \tilde{a}\,,\, \hat{B}\tilde{a}\rangle
-
\langle \tilde{a}_{+}\,,\, 2(1+\tau_{x}x^{2})^{1/2}\tilde{a}_{+}\rangle
-\frac{G_{\varepsilon,\tau}}{2}\int |a|^{4}
\right|
\\
\leq C_{\ell_{V},G}\left[\varepsilon^{2+4\delta}+ \varepsilon \times
\varepsilon^{2\delta}
\langle \tilde{a}\,,\, \hat{B}\tilde{a}\rangle\right]\,.
\end{multline*}
With
$\mathcal{E}_{\varepsilon}(\psi)=\mathcal{E}_{\varepsilon,min}$
($=\mathcal{O}(\varepsilon^{2+2\delta})$), this gives
$$
(1-C\varepsilon)\varepsilon^{2\delta}\langle
\tilde{a}\,,\,\hat{B}\tilde{a}\rangle+
2\langle
\tilde{a}_{+}\,,\,(1+\tau_{x}x^{2})^{1/2}\tilde{a}_{+}\rangle+\frac{G_{\varepsilon,\tau}}{2}\int|a|^{4}
\leq
C\varepsilon^{2+4\delta}+\mathcal{E}_{\varepsilon,min}\,,
$$
where all the terms of the left-hand side are now non negative. We
deduce \eqref{eq.L2a+} and by bootstrapping
\begin{multline*}
\varepsilon^{2+2\delta}\tau_{x}\langle a_{-}\,,\,
\hat{H}_{-}a_{-}\rangle +\frac{G_{\varepsilon,\tau}}{2}\int |a|^{4}
\leq \varepsilon^{2\delta}\langle
\tilde{a}\,,\,\hat{B}\tilde{a}\rangle
+\frac{G_{\varepsilon,\tau}}{2}\int |\tilde{a}|^{4}
\\
\leq \mathcal{E}_{\varepsilon,min}+C_{\ell_{V},G}\varepsilon^{2+4\delta}\,.
\end{multline*}
Additionally,
$\int|a|^{4}=\int\left(|a_{-}|^{2}+|a_{+}|^{2}\right)^{2}\geq
\int|a_{-}|^{4}$ also gives
$$
0\leq \|a_{-}\|_{L^{2}}^{4}\mathcal{E}_{\tau,min}\leq \mathcal{E}_{\tau}(a_{-})
\leq
C_{\ell_{V},G}\frac{\varepsilon^{2\delta}}{\tau_{x}}+\frac{\mathcal{E}_{\varepsilon,min}}{\varepsilon^{2+2\delta}\tau_{x}}\leq \mathcal{E}_{\tau,min}+C'_{\ell_{V},G}\frac{\varepsilon^{2\delta}}{\tau_{x}}\,.
$$
With
$\|a_{-}\|^{2}_{L^{2}}=1-\|a_{+}\|^{2}=1+\mathcal{O}(\varepsilon^{2+2\delta})=1+\mathcal{O}(\frac{\varepsilon^{2\delta}}{\tau_{x}})$
and $\mathcal{E}_{\tau,min}=\mathcal{E}_{H,min}+\mathcal{O}(\tau_{x}^{2/3})$,
this finally leads to
$$
|\frac{\mathcal{E}_{\varepsilon,min}}{\varepsilon^{2+2\delta}\tau_{x}}
-\mathcal{E}_{\tau}(a_{-})|\leq C''_{\ell_{V},G}\frac{\varepsilon^{2\delta}}{\tau_{x}}\,.
$$
{\hfill $\qed${}
\begin{lemme}
\label{le.ellipt} Assume $B(q,p,\tau,\varepsilon)=\tau_{x}|p|^{2}+\varepsilon r$ with $r\in
S_{u}(\langle \sqrt{\tau_{x}}p\rangle, g_{\tau};
\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, take any $\chi\in
\mathcal{C}^{\infty}_{0}(\field{R})$ and set $\hat{B}=B(q,\varepsilon
D_{q},\tau,\varepsilon)$ and
$\hat{\chi}=\chi(\tau_{x}\varepsilon|D_{q}|^{2})$\,. Then
there exists $\varepsilon_{B,\chi}>0$ and
$C_{B,\chi}>0$ such that the estimates
\begin{eqnarray*}
&&
\|(1-\hat{\chi})u\|_{L^{2}}^{2}+\|\sqrt{\tau_{x}}|\varepsilon
D_{q}|(1-\hat{\chi})u\|_{L^{2}}^{2}\leq C_{B,\chi}\left[\langle u\,,\,
\hat{B}u\rangle
+ \varepsilon^{1+2\delta}\|u\|_{L^{2}}^{2}\right]\,,
\\
\text{and}&&
\|(1-\hat{\chi})u\|_{L^{2}}+\|\sqrt{\tau_{x}}|\varepsilon
D_{q}|(1-\hat{\chi})u\|_{L^{2}}\leq C_{B,\chi}\left[
\|\hat{B}u\|_{L^{2}}
+ \varepsilon^{1+2\delta}\|u\|_{L^{2}}\right]\,,
\end{eqnarray*}
hold uniformly w.r.t $\delta\in (0,\delta_{0}]$ and $\varepsilon\in (0,\varepsilon_{B,\chi})$\,.
\end{lemme}
\noindent{\bf Proof: \ }{}
For $\chi_{0}\in \mathcal{C}^{\infty}_{0}(\field{R})$ such that
$\chi_{0}\geq 0$ and $\chi_{0}\equiv 1$ around $0$, the symbol
$\chi_{0}(\tau_{x}p^{2})+B$ is an elliptic symbol in $S_{u}(\langle
\sqrt{\tau_{x}}p\rangle^{2},g_{\tau}:\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,. For
$\varepsilon>0$ small enough according to $(\chi_{0},B)$, its quantization
$\hat{\chi}_{0}+\hat{B}$ is
invertible and its inverse belongs to $OpS_{u}(\langle
\sqrt{\tau_{x}}p\rangle^{-2},g_{\tau}, \mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,. Next we
notice
that in
$$
(1-\hat{\chi})=(1-\hat\chi)\circ\left[\hat{\chi}_{0}+\hat{B}\right]^{-1}\circ
\hat{B}
+
(1-\hat\chi)\circ\left[\hat{\chi}_{0}+\hat{B}\right]^{-1}\circ \hat{\chi}_{0}\,,
$$
the last term belongs to $Op\mathcal{N}_{u,g_{\tau}}$ if $\chi_{0}\prec
\chi$\,.
All the estimates are consequences of
$$
(1-\hat{\chi})=(1-\hat\chi)\circ\left[\hat{\chi}_{0}+\hat{B}\right]^{-1}\circ
\hat{B} +\varepsilon^{1+2\delta}\rho\,,
$$
with $\rho\in S_{u}(\frac{1}{\langle \tau_{x}p\rangle^{\infty}},g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$\,.
{\hfill $\qed${}
\subsection{Adiabatic Euler-Lagrange equation}
\label{se.adELeq}
As suggested by Proposition~\ref{pr.appH2}, or the last steps in the
proof of Proposition~\ref{pr.redHR}, an accurate comparison of
minimizers requires some comparison of the Euler-Lagrange equations.
We check here that the $a_{-}$ component in
Proposition~\ref{pr.compen} solves approximately the Euler-Lagrange
equations for minimizers of $\mathcal{E}_{\tau}$\,. Here the bootstrap
argument is made in terms of operators instead of quadratic forms.
In order to get reliable results, we now assume
$$
\varepsilon^{2\delta}\leq c_{\ell_{V},G}\tau_{x}\,,
$$
with $c_{\ell_{V},G}$ chosen small enough.\\
With such an assumption we know:
\begin{itemize}
\item from \eqref{eq.compEmin} and
$|\mathcal{E}_{\tau,min}-\mathcal{E}_{H,min}|=\mathcal{O}(\tau_{x}^{2/3})$,
$$
C^{-1}\varepsilon^{2+2\delta}\tau_{x}\leq
\mathcal{E}_{\varepsilon,min}\leq C\varepsilon^{2+2\delta}\tau_{x}\,.
$$
\item When $\psi\in \mathrm{Argmin}~\mathcal{E}_{\varepsilon,min}$ is
written $\psi=\hat{U}
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-\frac{y}{2\sqrt{\tau_{x}}}}a_{+}
\end{pmatrix}
$\, the $L^{4}$-norm of $a$ is uniformly bounded, $\int|a|^{4}\leq
C_{\ell_{V},G}$,
according to \eqref{eq.L2a+}.
By applying Lemma~\ref{le.Taylor} this also gives $\int|\psi|^{4}\leq
C_{\ell_{V},G}$\,.
We also have $\|a_{+}\|_{L^{2}}^{2}\leq C_{\ell_{V},G}\varepsilon^{2+2\delta}\tau_{x}$\,.
\item The Lagrange multiplier $\lambda_{\psi}$,
associated with $\psi\in
\mathrm{Argmin}~\mathcal{E}_{\varepsilon,min}$, equals
$\mathcal{E}_{\varepsilon}(\psi)+ \frac{G_{\varepsilon,\tau}}{2}\int
|\psi|^{4}$\,. Thus it is of order $\mathcal{O}(\varepsilon^{2+2\delta}\tau_{x})$\,.
\end{itemize}
\begin{proposition}
\label{pr.adEL}
Under the same assumptions as in Proposition~\ref{pr.compen} and with
the above condition $\varepsilon^{2\delta}\leq
c_{\ell_{V},G}\tau_{x}$, write a
minimizer $\psi$ of $\mathcal{E}_{\varepsilon}$ in the form $\psi=\hat{U}
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$\,.
Then the component $a_{-}$ solves the equation
$$
\hat{H}_{-}a_{-}+G\int|a_{-}|^{2}a_{-}=\lambda_{\psi}a_{-}+r_{\varepsilon}\quad
\text{with}\quad \|r_{\varepsilon}\|_{L^{2}}\leq
C_{\ell_{V},G}(\varepsilon+\frac{\varepsilon^{2\delta}}{\tau_{x}})\,,
$$
while $\|a_{+}\|_{L^{2}}\leq
C_{\ell_{V},G}\varepsilon^{2+2\delta}\tau_{x}$\,.\\
The $L^{p}$-norms of $a$ and $\psi$ are uniformly bounded by $C_{\ell_{V},G}$
for $p\in\left[2,6\right]$\,.\\
Moreover if $\hat{B}$ is the operator defined in
\eqref{eq.defm1}-\eqref{eq.defm2}, $\tilde{a}=\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$ satisfies
\begin{equation}
\label{eq.H2a}
\|\hat{B}\tilde{a}\|_{L^{2}}\leq C_{\ell_{V},G}\varepsilon^{2}\tau_{x}\,.
\end{equation}
\end{proposition}
\begin{remarque}
The relation of $\|\hat{B}\tilde{a}\|_{L^{2}}$ with
$\|\tilde{a}\|_{H^{2}}$ is given by
$$
C^{-1}
\left[\varepsilon^{2}\tau_{x}\|D_{q}^{2}\tilde{a}\|+ \varepsilon^{2}\|\tilde{a}\|_{L^{2}}\right]
\leq
\|\hat{B}\tilde{a}\|_{L^{2}}+ \varepsilon^{2}\|\tilde{a}\|_{L^{2}} \leq C
\left[\varepsilon^{2}\tau_{x}\|D_{q}^{2}\tilde{a}\|+ \varepsilon^{2}\|\tilde{a}\|_{L^{2}}\right]\,.
$$
Note that the $L^{2}$ remainder terms have a factor $\varepsilon^{2}$
and not $\varepsilon^{2}\tau_{x}$\,.
\end{remarque}
\noindent{\bf Proof: \ }{}
Playing with the Euler-Lagrange equation for $\psi$,
we shall first prove \eqref{eq.H2a} by using the same argument as we did
for $\langle \tilde{a}\,,\, \hat{B}\tilde{a}\rangle$ in the variational proof
of Proposition~\ref{pr.compen} and then use it in order
to estimate $\|r\|_{L^{2}}$\,.
The Euler-Lagrange equation for $\psi$
$$
H_{Lin}\psi+ G_{_{\varepsilon,\tau}}|\psi|^{2}\psi=\lambda_{\psi}\psi
$$
becomes
$$
\hat{U}^{*}H_{Lin}\hat{U}\tilde{a}+G_{\varepsilon, \tau}\hat{U}^{*}\left(|\psi|^{2}\psi\right)=\lambda_{\psi} \tilde{a}\,.
$$
Remember that Born-Oppenheimer Hamiltonian is given by
\begin{eqnarray*}
\varepsilon^{2+2\delta}\tau_{x}H_{BO}&=&
\varepsilon^{2+2\delta}\tau_{x}\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}\hat{H}_{+}e^{-i\frac{y}{2\sqrt{\tau_{x}}}}&0\\
0&e^{-i\frac{y}{2\sqrt{\tau_{x}}}}\hat{H}_{-}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}
\end{pmatrix}\,,
\\
&=&\varepsilon^{2\delta}\hat{B}+
\begin{pmatrix}
2\sqrt{1+\tau_{x}x^{2}}\\0
\end{pmatrix}\,,
\end{eqnarray*}
by using the notations of \eqref{eq.defm1}-\eqref{eq.defm2}.
Let us consider first the linear part after decomposing $\tilde{a}$
into $\tilde{a}=\hat{\chi}\tilde{a}+(1-\hat{\chi})\tilde{a}$, where
the kinetic energy cut-off operator
$\hat{\chi}=\chi(\tau_{x} |\varepsilon D_{q}|^{2})$ has been
introduced in Proposition~\ref{pr.adiabexpl2}:
\begin{multline*}
\hat{U}^{*}H_{Lin}\hat{U}\tilde{a}=\underbrace{\hat{U}^{*}H_{Lin}\hat{U}\hat{\chi}\tilde{a}}_{(I)}
\\
+\underbrace{(\hat{U}^{*}-\hat{U}_{0}^{*})H_{Lin}\hat{U}(1-\hat{\chi})\tilde{a}
+ \hat{U}_{0}^{*}H_{Lin}(\hat{U}-\hat{U}_{0})(1-\hat{\chi})\tilde{a}
}_{(II)}
\\
+ \underbrace{\hat{U}_{0}^{*}H_{Lin}\hat{U}_{0}(1-\hat{\chi})\tilde{a}}_{(III)}\,.
\end{multline*}
The three terms are treated by reconsidering the computations done for
Proposition~\ref{pr.adiabexpl2}. By inserting a cut-off
$\hat{\chi_{1}}$\,, $\chi\prec\chi_{1}$\,, in
$\chi_{1}\hat{U}^{*}H_{Lin}\hat{U}\hat{\chi}\tilde{a}$ with
$(1-\hat{\chi}_{1})\hat{U}^{*}H_{Lin}\hat{U}\hat{\chi}\in
Op\mathcal{N}_{u,g_{\tau}}$, we get
$$
(I)=\varepsilon^{2+2\delta}\tau_{x}H_{BO}\hat{\chi}\tilde{a}
+\mathcal{O}(\varepsilon^{2+4\delta})\quad\text{in}\, L^{2}(\field{R}^{2})\,.
$$
With $\varepsilon^{-1-2\delta}(\hat{U}-\hat{U_{0}})\in
OpS_{u}(\frac{1}{\langle \sqrt{\tau_{x}}p\rangle^{\infty}},
g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$ and by using Lemma~\ref{le.ellipt},
the second term is estimated by
$$
\|(II)\|_{L^{2}}\leq
C\varepsilon^{1+2\delta}\|(1-\hat{\chi})\tilde{a}\|_{L^{2}}
\leq
C'\left[\varepsilon\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}+ \varepsilon^{2+4\delta}\right]\,.
$$
We write the third term as
$$
(III)=\varepsilon^{2+2\delta}\tau_{x}H_{BO}(1-\hat{\chi})\tilde{a} +
\varepsilon^{2+2\delta}\tau_{x}\mathcal{D}_{kin}(1-\hat{\chi})\tilde{a}+\mathcal{D}_{pot}(1-\hat{\chi})\tilde{a}\,,
$$
where $\mathcal{D}_{kin}$ and $\mathcal{D}_{pot}$ are defined by
\eqref{eq.Dkin}-\eqref{eq.Dpot}. Following the arguments given in the
proof of Proposition~\ref{pr.adiabexpl2} after these definitions, we
get
\begin{eqnarray*}
\|\mathcal{D}_{pot}(1-\hat{\chi})\tilde{a}\|_{L^{2}}
&\leq &C\varepsilon^{1+2\delta}\|(1-\hat{\chi})\tilde{a}\|_{L^{2}}
\leq
\\
&\leq&
C'\left[\varepsilon\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}+ \varepsilon^{2+4\delta}\right]\,,
\\
\|\mathcal{D}_{kin}(1-\hat{\chi})\tilde{a}\|_{L^{2}}
&\leq &
C\left[\varepsilon^{1+4\delta}\|(1-\hat{\chi})\tilde{a}\|_{L^{2}}+
\varepsilon^{1+2\delta}\|(\varepsilon
\sqrt{\tau_{x}}|D_{q}|)(1-\hat{\chi})\tilde{a}\|_{L^{2}}\right]
\\
&\leq&
C'\left[\varepsilon\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}+ \varepsilon^{2+4\delta}\right]\,.
\end{eqnarray*}
Hence the Euler-Lagrange equation can be written
\begin{eqnarray*}
&&
\varepsilon^{2+2\delta}\tau_{x}H_{BO}\tilde{a}=\lambda_{\psi}\tilde{a}-\varepsilon^{2+2\delta}\tau_{x}G
\hat{U}^{*}\left(|\psi|^{2}\psi\right)+r
\\
\text{with}
&&
\|r\|_{L^{2}}\leq C
\left[
\varepsilon\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}
+\varepsilon^{2+4\delta}\right]\,.
\end{eqnarray*}
From Proposition~\ref{pr.existEeps}, with $\varepsilon^{2\delta}\leq
c\tau_{x}$, we know $\|\psi\|_{L^{4}}\leq C$\,. It can be transformed
into $\|\tilde{a}\|_{L^{4}}=\|\hat{U}^{*}\psi\|_{L^{4}}\leq C$ with
$$
\||\psi|^{2}\psi\|_{L^{2}}\leq C\|\psi\|_{L^{6}}^{3}\leq C'\|\tilde{a}\|_{L^{6}}^{3}\,,
$$
by applying
Lemma~\ref{le.Taylor}, adapted for the metric $g_{\tau}$ like
in the proof Proposition~\ref{pr.L4err}.
We start from the interpolation inequality
$$
\|f\|_{L^{6}}\leq
C\|D_{q}^{2}f\|_{L^{2}}^{\frac{1}{9}}\|f\|_{L^{4}}^{\frac{8}{9}}
\leq \frac{C}{(\tau_{x}\varepsilon)^{\frac{2}{9}}}
\|\tau_{x}^{2}\varepsilon^{2}D_{q}^{2}f\|_{L^{2}}^{\frac{1}{9}}\|f\|_{L^{4}}^{\frac{8}{9}}.
$$
By introducing the operator $\tilde{B}=B(\tau_{x}^{-1/2}\varepsilon q,
\tau_{x}^{\frac{1}{2}}D_{q},\tau,\varepsilon)$ with
\begin{multline*}
B_{\pm}(\tau_{x}^{-1/2} q, \tau_{x}^{1/2}D_{q},\tau,\varepsilon)=(\tau_{x}\varepsilon)^{2}D_{q}^{2}\pm
\frac{\varepsilon}{2}\left(\frac{x}{\sqrt{1+x^{2}}}-\sqrt{\tau_{x}}\right)(\tau_{x}\varepsilon)D_{y}
\\
+\frac{\varepsilon^{2}}{4}\left[\left(\frac{x}{\sqrt{1+x^{2}}}-\sqrt{\tau_{x}}\right)^{2}+\frac{4v}{\ell_{V}^{2}}\right]\,,
\end{multline*}
it becomes
$$
\|f\|_{L^{6}}\leq\frac{C_{\ell_{V}}}{(\tau_{x}\varepsilon)^{\frac{2}{9}}}
\left[\|\tilde{B}f\|_{L^{2}}^{\frac{1}{9}}\|f\|_{L^{4}}^{\frac{8}{9}}
+ \varepsilon^{\frac{2}{9}}\|f\|_{L^{2}}^{\frac{1}{9}}\|f\|_{L^{4}}^{\frac{8}{9}}\right]\,.
$$
The above relation with $f=\tau_{x}^{-\frac{1}{2}}\tilde{a}(\tau_{x}^{-\frac{1}{2}}.)$ which satisfies
$$
\|f\|_{L^{6}}=\tau_{x}^{-\frac{1}{3}}\|\tilde{a}\|_{L^{6}}\quad,\quad
\|f\|_{L^{4}}=\tau_{x}^{-\frac{1}{2}}\|\tilde{a}\|_{L^{4}}
\quad\text{and}\quad
\|\tilde{B}f\|_{L^{2}}=\|\hat{B}\tilde{a}\|_{L^{2}}\,
$$
leads to
\begin{eqnarray*}
\tau_{x}^{-\frac{1}{3}}\|\tilde{a}\|_{L^{6}}
&\leq&
\frac{C_{\ell_{V}}}{(\tau_{x}\varepsilon)^{\frac{2}{9}}}
\left[\|\hat{B}\tilde{a}\|_{L^{2}}^{\frac{1}{9}}\tau_{x}^{-\frac{4}{9}}\|\tilde{a}\|_{L^{4}}^{\frac{8}{9}}
+
\varepsilon^{\frac{2}{9}}\|\tilde{a}\|_{L^{2}}^{\frac{1}{9}}\tau_{x}^{-\frac{4}{9}}\|\tilde{a}\|_{L^{4}}^{\frac{8}{9}}\right]
\\
&\leq &
C_{\ell_{V},G}\tau_{x}^{-\frac{1}{3}}
\left[\varepsilon^{-\frac{2}{9}}\|\hat{B}\tilde{a}\|_{L^{2}}^{\frac{1}{9}}+1\right]\,,
\end{eqnarray*}
and finally
$$
G_{\varepsilon,\tau}\||\psi|^{2}\psi\|_{L^{2}}\leq
G\varepsilon^{2+2\delta}\tau_{x}\|\psi\|_{L^{6}}^{3}
\leq
C_{\ell_{V},G}\left[\varepsilon^{4/3}\tau_{x}\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}+
\varepsilon^{2+2\delta}\tau_{x}\right]
$$
With $|\lambda_{\psi}|=\mathcal{O}(\varepsilon^{2+2\delta}\tau_{x})$, we obtain
$$
\|\varepsilon^{2+2\delta}\tau_{x}H_{BO}\tilde{a}\|_{L^{2}}
\leq
C
\left[\varepsilon\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}+\varepsilon^{2+4\delta}+\varepsilon^{2+2\delta}\tau_{x}\right]\,,
$$
and recall
$$
\varepsilon^{2+2\delta\tau_{x}}H_{BO}=
\begin{pmatrix}
\varepsilon^{2\delta}\hat{B}_{+}+2\sqrt{1+\tau_{x}x^{2}}\\
\varepsilon^{2\delta}\hat{B}_{-}
\end{pmatrix}\,.
$$
According to Lemma~\ref{le.ellipt2} below
$$
\|\varepsilon^{2\delta}\hat{B}_{+}\tilde{a}_{+}\|_{L^{2}}+
\|2\sqrt{1+\tau_{x}x^{2}}\tilde{a}_{+}\|_{L^{2}}
\leq
C\|(\varepsilon^{2\delta}\hat{B}_{+}+2\sqrt{1+\tau_{x}x^{2}})\tilde{a}_{+}\|_{L^{2}}\,.
$$
We deduce
$$
\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}}\leq C\varepsilon^{2+2\delta}\tau_{x}\,.
$$
Plugging this result into the estimate of the remainder $r$ gives
$$
\varepsilon^{2+2\delta}\tau_{x}H_{BO}\tilde{a}=\lambda_{\tau}\tilde{a}+
G_{\varepsilon,\tau}\hat{U}^{*}(|\psi|^{2}\psi)+ \mathcal{O}(\varepsilon^{3+2\delta}\tau_{x}+\varepsilon^{2+4\delta})\,.
$$
Consider now more carefully the nonlinear term
\begin{multline*}
\hat{U}^{*}\left(|\psi|^{2}\psi\right)
=
\underbrace{
\hat{U}_{0}^{*}\left[|\hat{U}_{0}\tilde{a}|^{2}(\hat{U}_{0}\tilde{a})\right]}_{(1)}
+
\underbrace{
(\hat{U}^{*}-\hat{U}_{0}^{*})\left[|\hat{U}_{0}\tilde{a}|^{2}
(\hat{U}_{0}\tilde{a})\right]}_{(2)}
\\
+
\underbrace{
\hat{U}^{*}\left[|\hat{U}\tilde{a}|^{2}
(\hat{U}\tilde{a})
-
|\hat{U}_{0}\tilde{a}|^{2}
(\hat{U}_{0}\tilde{a})\right]
}_{(3)}\,.
\end{multline*}
By differentiating the relation $\int |f|^{4}=\int
|\hat{U}_{0}f|^{4}$ w.r.t $f$, the first term equals
$$
(1)=|\tilde{a}|^{4}\tilde{a}\,.
$$
By semiclassical calculus in the metric $g_{\tau}$, the operator
$ \sqrt{\tau_{x}}\varepsilon D_{q}\hat{U}_{0}$ equals
$$
\sqrt{\tau_{x}}\varepsilon D_{q}\hat{U}_{0}
=\hat{U}_{0}\sqrt{\tau_{x}}\varepsilon D_{q} +
\left[\sqrt{\tau_{x}}\varepsilon D_{q}, \hat{U}_{0}\right]\psi
=\hat{U}_{0}\sqrt{\tau_{x}}\varepsilon D_{q}+
\mathcal{O}(\varepsilon)\,,
$$
where the remainder estimate holds in $\mathcal{L}(L^{2}(\field{R}^{2};\field{C}} \newcommand{\nz}{\field{N}^{2}))$\,.
We have already proved $\|\tilde{a}\|_{L^{6}}\leq C$ and
Lemma~\ref{le.Taylor} leads again to
$$
\|\hat{U}\tilde{a}\|_{L^{6}}+
\|\hat{U}_{0}\tilde{a}\|_{L^{6}}\leq
C\,.
$$
With $\varepsilon^{-1-2\delta}(\hat{U}-\hat{U}_{0})\in
OpS_{u}(1,g_{\tau};\mathcal{M}_{2}(\field{C}} \newcommand{\nz}{\field{N}))$, this gives
$$
\|(2)\|\leq C\varepsilon^{1+2\delta}\,.
$$
For the third term, we use
\begin{eqnarray*}
\|\left[|\hat{U}\tilde{a}|^{2}
(\hat{U}\tilde{a})
-
|\hat{U}_{0}\tilde{a}|^{2}
(\hat{U}_{0}\tilde{a})\right]\|_{L^{2}}
&&
\leq
C_{0}\|(\hat{U}-\hat{U}_{0})\tilde{a}\|_{L^{6}}
\left[\|\hat{U}\tilde{a}\|_{L^{6}}^{2}+\|\hat{U}_{0}\tilde{a}\|_{L^{6}}^{2}\right]
\\
&&\leq C\varepsilon^{1+2\delta}\,.
\end{eqnarray*}
We have proved
$$
G_{\tau,\varepsilon}\hat{U}^{*}(|\psi|^{2}\psi)-G_{\tau,\varepsilon}|\tilde{a}|^{2}\tilde{a}
=\mathcal{O}(\varepsilon^{3+4\delta}\tau_{x})\,,
$$
which is even better than the estimate for the linear part.\\
For the $\tilde{a}_{-}$ component, we get
$$
H_{BO}
\begin{pmatrix}
0\\\tilde{a}_{-}
\end{pmatrix}
+G(|\tilde{a}_{-}|^{2}+|\tilde{a}_{+}|^{2})\tilde{a}_{-}
=\frac{\lambda_{\psi}}{\varepsilon^{2+2\delta}\tau_{x}}\tilde{a}_{-}+\mathcal{O}(\varepsilon+
\frac{\varepsilon^{2\delta}}{\tau_{x}})\,,
$$
and it remains to estimate the term
$|\tilde{a}_{+}|^{2}\tilde{a}_{-}$\,.
The first line of the system may be written
$$
\varepsilon^{2\delta}\hat{B}_{+}\tilde{a}_{+}+
2\sqrt{1+\tau_{x}x^{2}}\tilde{a_{+}}
+
G_{\varepsilon,\tau}\left(|\tilde{a}_{+}|^{2}+|\tilde{a}_{-}|^{2}\right)
\tilde{a}_{+}
=\lambda_{\psi}\tilde{a}_{+}+
\mathcal{O}(\varepsilon^{3+2\delta}\tau_{x}+ \varepsilon^{2+2\delta}\tau_{x})\,.
$$
Taking the scalar product with $\tilde{a}_{+}$, with
$\lambda_{\tau}=\mathcal{O}(\varepsilon^{2+2\delta}\tau_{x})$ and
$\hat{B}_{+}\geq 0$, gives
$$
\int |\tilde{a}_{+}|^{4}\leq
C[\|\tilde{a}_{+}\|_{L^{2}}^{2}+(\varepsilon+
\frac{\varepsilon^{2\delta}}{\tau_{x}})\|a_{+}\|_{L^{2}}]
\leq C'\varepsilon^{2+2\delta}\tau_{x}\,.
$$
Then the same argument as in the estimate of
$\|\tilde{a}\|_{L^{6}}$ gives
\begin{eqnarray*}
\|\tilde{a}_{+}\|_{L^{6}}
&\leq&
C\left[\varepsilon^{-\frac{2}{9}}\|\hat{B}_{+}\tilde{a}_{+}\|_{L^{2}}^{\frac{1}{9}}\|\tilde{a}_{+}\|_{L^{4}}^{\frac{8}{9}}
+
\|\tilde{a}_{+}\|_{L^{2}}^{\frac{1}{9}}\|\tilde{a}_{+}\|_{L^{4}}^{\frac{8}{9}}\right]
\\
&\leq &
C'
\left[(\varepsilon^{-2-2\delta}\|\varepsilon^{2\delta}\hat{B}\tilde{a}\|_{L^{2}})^{\frac{1}{9}}
+
(\varepsilon^{2+2\delta}\tau_{x})^{\frac{1}{18}}
\right](\varepsilon^{2+2\delta}\tau_{x})^{\frac{2}{9}}\\
&\leq&
C''\tau_{x}^{\frac{1}{9}}(\varepsilon^{2+2\delta}\tau_{x})^{\frac{2}{9}}
\leq C'' \varepsilon^{\frac{4+4\delta}{9}}\tau_{x}^{\frac{1}{3}}\,.
\end{eqnarray*}
Again with $\|\tilde{a}\|_{L^{6}}\leq C$, we deduce
$$
\||\tilde{a}_{+}|^{2}\tilde{a}_{-}\|_{L^{2}}\leq
C_{\ell_{V},G}\varepsilon^{\frac{8+8\delta}{3}}\tau_{x}^{\frac{2}{3}}\leq
C_{\ell_{V},G}'\varepsilon\,.
$$
The final result is just a transcription in terms of $a$\,.
{\hfill $\qed${}
\begin{lemme}
\label{le.ellipt2}
Let $B_{+}$ be the symbol
$$
B_{+}(q,p,\tau_{x},\varepsilon)=\tau_{x}p_{x}^{2}+(\sqrt{\tau_{x}}p_{y}+\frac{\varepsilon}{2}(\frac{\sqrt{\tau_{x}}x}{\sqrt{1+\tau_{x}x^{2}}}-\sqrt{\tau_{x}}))^{2}
+\frac{\varepsilon^{2}v(\sqrt{\tau_{x}}.)}{\ell_{V}^{2}}
$$
introduced in \eqref{eq.defm2}. By setting $\hat{B}_{+}=B_{+}(q,\varepsilon
D_{q},\tau_{x},\varepsilon)$, the operator
$A=\varepsilon^{2\delta}\hat{B}_{+}+2\sqrt{1+\tau_{x}x^{2}}$
is self-adjoint with $D(A)=\left\{u\in L^{2}(\field{R}^{2}),\, Au\in
L^{2}(\field{R}^{2})\right\}$
as soon as $\varepsilon\leq \varepsilon_{0}$, with $\varepsilon_{0}$
independent of $\tau_{x}$\,. Moreover the inequality
$$
\forall u\in D(A),\quad\|\varepsilon^{2\delta}\hat{B}_{+}u\|_{L^{2}}+\|2\sqrt{1+\tau_{x}x^{2}}u\|_{L^{2}}
\leq C \|Au\|_{L^{2}}
$$
holds with a constant $C$ independent of $(\varepsilon,\tau_{x})$\,.
\end{lemme}
\noindent{\bf Proof: \ }{}
The operator $A$ can be written
$$
A=a_{0}(q,\varepsilon^{1+\delta}D_{q},\tau_{x})+
\varepsilon^{1+\delta}a_{1}(q,\varepsilon^{1+\delta}D_{q},\tau_{x})+\varepsilon^{2+2\delta}a_{2}(q,\tau_{x})\,,
$$
with
$a_{k}\in S_{u}(\langle \sqrt{\tau_{x}}p\rangle^{2-k}+\langle
\sqrt{\tau_{x}}x\rangle^{(1-k)_{+}}, g_{\tau})$ and
$$
a_{0}(q,p,\tau_{x})=p^{2}+2\sqrt{1+\tau_{x}x^{2}}\,.
$$
Therefore the operator $A$ is elliptic in $OpS_{u}(\langle \sqrt{\tau_{x}}p\rangle^{2}+\langle
\sqrt{\tau_{x}}x\rangle, g_{\tau})$ and the result about the domain
follows with
$$
\|\varepsilon^{2\delta}\hat{B}_{+}u\|_{L^{2}}+\|2\sqrt{1+\tau_{x}x^{2}}u\|_{L^{2}}
\leq C (\|Au\|_{L^{2}}+\|u\|_{L^{2}})
$$
for all $u\in D(A)$\,. We conclude with
$$
2\|u\|_{L^{2}}^{2}\leq \langle u\,,\, Au\rangle\leq \|u\|_{L^{2}}\|Au\|_{L^{2}}
$$
due to $\hat{B}_{+}\geq 0$\,.
{\hfill $\qed${}
\subsection{End of the proof of Theorem \ref{th.mintot}}
\label{se.mintot}
Assume $G,\ell_{V}>0$ be fixed and $\delta\in
(0,\delta_{0}]$\,. Although we dropped $\delta_{0}$ in our notations,
all the constants in the previous inequalities depend on
$(G,\ell_{V},\delta_{0})$\,.
We assume now $\tau_{x}\leq \tau_{\ell_{V},G,\delta_{0}}$,
$\varepsilon\leq \varepsilon_{\ell_{V},G,\delta_{0}}$ and
$$
\varepsilon^{2\delta}\leq \tau_{x}^{\frac{5}{3}}\,.
$$
Proposition~\ref{pr.compen} says
$$
|\mathcal{E}_{\varepsilon,min}-\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau,min}|\leq
C\varepsilon^{2+4\delta}\leq C'\varepsilon^{2+2\delta}\tau_{x}^{\frac{5}{3}}
$$
and we recall
$$
|\mathcal{E}_{\tau,min}-\mathcal{E}_{H,min}|\leq C\tau_{x}^{\frac{2}{3}}
$$
by Proposition~\ref{pr.redHR}.\\
When $\psi=\hat{U}
\begin{pmatrix}
e^{i\frac{y}{2\sqrt{\tau_{x}}}}a_{+}\\
e^{-i\frac{y}{2\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$\,, Proposition~\ref{pr.adEL} says
$$
\|a_{+}\|_{L^{2}}\leq C\varepsilon^{2+2\delta}\tau_{x}
\quad,\quad
\|a\|_{H^{2}}\leq \frac{C}{\tau_{x}}
\quad\text{and}\quad
\|a\|_{L^{4}}+\|a\|_{L^{6}}\leq C\,,
$$
while $a_{-}$ solves the approximate Euler-Lagrange equation
$$
\hat{H}_{-}a_{-}+G\int|a_{-}|^{2}a_{-}=\lambda_{\psi}a_{-}+r_{\varepsilon}\quad
\text{with}\quad \|r_{\varepsilon}\|_{L^{2}}\leq
C_{\ell_{V},G}(\varepsilon+\frac{\varepsilon^{2\delta}}{\tau_{x}})\,.
$$
For $u=\|a_{-}\|_{L^{2}}^{-1}a_{-}$ the energy $\mathcal{E}_{\tau}(u)$ satisfy
$$
|\mathcal{E}_{\tau}(u)-\mathcal{E}_{\tau,min}|
\leq C\tau_{x}^{\frac{2}{3}}
$$
and the above equation becomes
$$
\hat{H}_{-}u+G\int|u|^{2}u=\lambda_{\psi}u+
\mathcal{O}(\varepsilon+\frac{\varepsilon^{2\delta}}{\tau_{x}})\,.
$$
We conclude by referring to Proposition~\ref{pr.appH2} applied to $u$
and then renormalizing for $a_{-}$: For
$\chi=(\chi_{1},\chi_{2})$ with $\chi_{1}\in
\mathcal{C}^{\infty}_{0}(\field{R}^{2})$, $\chi_{1}^{2}+\chi_{2}^{2}\equiv
1$, $\chi_{1}=1$ in a neighborhood of $0$
\begin{eqnarray*}
&& \|\chi_{2}(\tau_{x}^{\frac{1}{9}}.)a_{-}\|_{L^{2}}\leq
C\tau_{x}^{\frac{1}{3}}\\
&&
|\mathcal{E}_{\tau}(\chi_{1}(\tau_{x}^{\frac{1}{9}}a_{-})-\mathcal{E}_{H,min})|\leq
C\tau_{x}^{\frac{2}{3}}
\\
&&
d_{\mathcal{H}_{2}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\mathrm{Argmin}~{\mathcal{E}_{H}})\leq
C(\tau_{x}^{\frac{2\nu_{\ell_{V},G}}{3}}+\varepsilon)\,,\quad
\nu_{\ell_{V},G}\in (0,\frac{1}{2}]\,.
\end{eqnarray*}
For the $L^{\infty}$-estimates of $a_{+}$ and $d_{L^{\infty}}(\chi_{1}(\tau_{x}^{\frac{1}{9}}.)a_{-},
\mathrm{Argmin}~{\mathcal{E}_{H}})$, we simply use the interpolation
inequality
$$
\|u\|_{L^{\infty}}\leq C\|u\|_{L^{2}}^{\frac{1}{2}}\|\Delta
u\|_{L^{2}}^{\frac{1}{2}}\leq C\|u\|_{L^{2}}^{\frac{1}{2}}\|u\|_{H^{2}}^{\frac{1}{2}}\,,
$$
valid for any $u\in H^{2}(\field{R}^{2})$ (write
$u(0)=\int_{\field{R}^{2}}\hat{u}(\xi)\frac{d\xi}{(2\pi)^{2}}$, cut the
integral according to $|\xi|\lessgtr R$, estimate both term by
Cauchy-Schwartz with $\hat{u}=\frac{1}{|\xi|^{2}}(|\xi|^{2}\hat{u})$
when $|\xi|\geq R$, and then optimize w.r.t $R$).
\section{Additional comments}
\label{se.complements}
We briefly discuss and sketch how our analysis could be adapted to
other problems. No definite statement
is given. Complete proofs require additional work, which may be
done in the future.
\subsection{About the smallness condition of $\varepsilon$ w.r.t $\tau_{x}$}
\label{se.epstau}
In our main result, Theorem~\ref{th.mintot}, condition \eqref{eq.condtaueps} is used, namely
$$
\varepsilon^{2\delta}\leq \frac{\tau_{x}^{\frac{5}{3}}}{C_{\delta}}\,.
$$
One may wonder whether such a condition is necessary in order to
compare the minimization problems for $\mathcal{E}_{\varepsilon}$ and
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{H}$\,.
When comparing the minimal energies in
Proposition~\ref{pr.compen}, we found
$$
|\mathcal{E}_{\varepsilon,min}-\varepsilon^{2+2\delta}\mathcal{E}_{H,min}|\leq C\varepsilon^{2+4\delta}\,,
$$
while we know that
$\mathcal{E}_{H,min}=\mathcal{E}_{H,min}(\ell_{V},G)$ is a positive
number independent of $\tau_{x}$ and $\varepsilon$\,. Hence it seems
natural to say that $\varepsilon^{2\delta}\ll \tau_{x}$, at least, is
required to ensure that
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{H,min}$ is a good
approximation of $\mathcal{E}_{\varepsilon,min}$\,. The error is made of three parts:
\begin{itemize}
\item the error term for the Born-Oppenheimer approximation in the
low-frequency range given in Theorem~\ref{th.BornOp};
\item the error term coming from the truncated high frequency part;
\item the non linear term.
\end{itemize}
The non linear term is
$\frac{G\varepsilon^{2+2\delta}\tau_{x}}{2}\int|\psi|^{4}$, so that a
small error in $\|\psi\|_{L^{4}}^{4}$ will give a negligible term w.r.t
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{H,min}$\,.
The question is thus mainly about the linear problem. If one looks
more carefully at the error term of Theorem~\ref{th.BornOp}, it is
made of the term
\begin{equation}
\label{eq.correc}
-\varepsilon^{2}\frac{(\partial_{p_{k}}f_{\varepsilon})(\partial_{p_{\ell}}f_{\varepsilon})}{E_{+}-E_{-}}
\overline{X}_{k}X_{\ell}\,,
\end{equation}
according to Proposition~\ref{pr.hameff}, and of terms coming from the
third order term of Moyal products.
The function $f_{\varepsilon}$ is in our case
$f_{\varepsilon}(p)=\varepsilon^{2\delta}\tau_{x}p^{2}\gamma(\tau_{x}p^{2})$,
$p=(p_{x},p_{y})$, $k,\ell\in \left\{x,y\right\}$\,, and the factors
$X_{x}$ and $X_{y}$ computed in the proof of
Proposition~\ref{pr.adiabexpl1} are at most of order
$\frac{1}{\sqrt{\tau_{x}}}$\,. Hence the quantity \eqref{eq.correc} is
an $\mathcal{O}(\varepsilon^{2+4\delta}\tau_{x})$ which is again
negligible w.r.t
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{H,min}$\,.
By considering the higher order terms in the Moyal product, the
fast oscillating part of the symbol w.r.t $y$, at the frequency
$\frac{1}{\sqrt{\tau_{x}}}$, deteriorates the estimates:
although there are compensations with the slow variations w.r.t $p_{y}$,
always multiplied by $\sqrt{\tau_{x}}$, only an $\varepsilon^{k}$ factor without
$\tau_{x}$ appears in the $k$-th order term.\\
Hence, computing the higher order terms,
at least up to order $3$, in the adiabatic approximation
and then considering the question
of the high-frequency truncation, is a way to
understand whether the
smallness of $\varepsilon$ w.r.t $\tau_{x}$ is necessary.
\subsection{Anisotropic nonlinearity}
\label{se.aniso}
Our work assumes an isotropic nonlinearity. A more general nonlinear term
would be
$$
\frac{G\varepsilon^{2+2\delta}\tau_{x}}{2}\int
\alpha_{1}|\psi_{1}|^{4}+2\alpha_{12}|\psi_{1}|^{2}|\psi_{2}|^{2}+
\alpha_{2}|\psi_{2}|^{4}~dxdy\,,\quad
\psi=
\begin{pmatrix}
\psi_{1}\\\psi_{2}
\end{pmatrix}\,.
$$ Our case is $\alpha_{1}=\alpha_{2}=\alpha_{12}=1$.
Let
$\psi=\hat{U}\phi$, with
the unitary transform
$\hat{U}=\hat{U_{0}}+\varepsilon^{1+2\delta}\hat{R}$, (or conversely
$\phi=\hat{U}^{*}\psi$). Then the same arguments as in
Subsection~\ref{se.nonlin} will lead to
\begin{multline*}
\int \alpha_{1}|\psi_{1}|^{4}+2\alpha_{12}|\psi_{1}|^{2}|\psi_{2}|^{2}
+ \alpha_{2}|\psi_{2}|^{4}~dxdy
\\
=
\left (\int \alpha_{1}|\psi_{1}^{0}|^{4}+2\alpha_{12}|\psi_{1}^{0}|^{2}|\psi_{2}^{0}|^{2}
+ \alpha_{2}|\psi_{2}^{0}|^{4}~dxdy\right )
(1+
\mathcal{O}(\varepsilon^{1+2\delta}))\,,
\end{multline*}
after setting $\psi^{0}(q)=u_{0}(q)\phi(q)$ at every $q=(x,y)$\,.
In our case
$$
u_{0}(x,y)=
\begin{pmatrix}
C&Se^{i\varphi}\\
Se^{-i\varphi}&-C
\end{pmatrix}
\quad, \quad C=\cos\left(\frac{\theta}{2}\right)
\,,\,
S=\sin\left(\frac{\theta}{2}\right)\,,
$$
with $\theta=\underline{\theta}(\sqrt{\tau_{x}}x)\,,\, \varphi=\frac{y}{\sqrt{\tau_{x}}}$\,.
The point-wise identities
\begin{eqnarray*}
&&
|\psi_{1}^{0}|^{2}+|\psi_{2}^{0}|^{2}
= |\phi_{1}|^{2}+|\phi_{2}|^{2}
\\
&&
|\psi_{1}^{0}|^{2}-|\psi_{2}^{0}|^{2}
=
\cos(\theta)\left(|\phi_{1}|^{2}-|\phi_{2}|^{2}\right)
+2\sin(\theta)\Real(\overline{\phi_{1}e^{i\varphi}}\phi_{2})\,,
\end{eqnarray*}
lead to
\begin{eqnarray*}
&&
\alpha_{1}|\psi_{1}^{0}|^4+2\alpha_{12}|\psi_{1}^{0}|^{2}|\psi_{2}^{0}|^{2}+\alpha_{2}|\psi_{2}^{0}|^{4}
=
\frac{\alpha_{1}+2\alpha_{12}+\alpha_{2}}{4}\left(|\phi_{1}|^{2}+|\phi_{2}|^{2}\right)^{2}
\\
&&\;+ \frac{\alpha_{1}-\alpha_{2}}{2}\left(|\phi_{1}|^{2}+|\phi_{2}|^{2}\right)
\left(\cos(\theta)(|\phi_{1}|^{2}-|\phi_{2}|^{2})+2\sin(\theta)
\Real(\overline{\phi_{1}e^{i\varphi}}
\phi_{2})\right)
\\
&&\;+
\frac{\alpha_{1}-2\alpha_{12}+\alpha_{2}}{4}
\left(\cos(\theta)(|\phi_{1}|^{2}-|\phi_{2}|^{2})+2\sin(\theta)
\Real(\overline{\phi_{1}e^{i\varphi}}
\phi_{2})\right)^{2}
\\
&&
\hspace{2cm}
=
\left[\alpha_{1}\cos^{4}(\frac{\theta}{2})+\alpha_{2}\sin^{4}(\frac{\theta}{2})+
\frac{\alpha_{12}}{2}\sin^{2}(\theta)\right]|\phi_{1}|^{4}
\\
&&
\hspace{3cm}
+
\left[\alpha_{1}\sin^{4}(\frac{\theta}{2})+\alpha_{2}\cos^{4}(\frac{\theta}{2})+
\frac{\alpha_{12}}{2}\sin^{2}(\theta)\right]|\phi_{2}|^{4}
\\
&&
\hspace{3cm}
+ \left[\frac{(\alpha_{1}+\alpha_{2})}{2}\sin^{2}(\theta)+
\alpha_{12}(1+\cos^{2}(\theta))\right]|\phi_{1}|^{2}|\phi_{2}|^{2}
\\
&&
\hspace{1cm}
+\left[(\alpha_{1}-\alpha_{2})+(\alpha_{1}-2\alpha_{12}+
\alpha_{2})\cos(\theta)\right]\sin(\theta)|\phi_{1}|^{2}
\Real(\overline{\phi_{1}e^{i\varphi}}\phi_{2})
\\
&&
\hspace{1cm} +\left[(\alpha_{1}-\alpha_{2})-(\alpha_{1}-2\alpha_{12}+
\alpha_{2})\cos(\theta)\right]\sin(\theta)|\phi_{2}|^{2}
\Real(\overline{\phi_{1}e^{i\varphi}}\phi_{2})
\\
&&
\hspace{3cm}
+(\alpha_{1}-2\alpha_{12}+\alpha_{2})\sin^{2}(\theta)\Real\left[(\overline{\phi_{2}e^{i\varphi}}\phi_{2})\right]\,.
\end{eqnarray*}
At least three points have to be adapted from the previous analysis:
\begin{description}
\item[1)] When we take a test function $\psi=\hat{U}
\begin{pmatrix}
0\\e^{-i\frac{y}{\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$
the energy $\mathcal{E}_{\varepsilon}(\psi)$ will be close to
$\varepsilon^{2+2\delta}\tau_{x}\mathcal{E}_{\tau}(a_{-})$, with
\begin{multline*}
\mathcal{E}_{\tau}(a_{-})=\langle a_{-}\,,\,\hat{H}_{-}a_{-}\rangle
\\
+\frac{G}{2}
\int
\left[\alpha_{1}\cos^{4}(\frac{\theta}{2})+\alpha_{2}\sin^{4}(\frac{\theta}{2})+
\frac{\alpha_{12}}{2}\sin^{2}(\theta)\right]|a_{-}|^{4}~dxdy\,,
\end{multline*}
and
$\cos(\theta)=\frac{\sqrt{\tau_{x}}x}{\sqrt{1+\tau_{x}x^{2}}}$\,. Hence
before taking the limit $\tau_{x}\to 0$ we have a position dependent
nonlinearity.
This will induce another error term when comparing with the
energy $\mathcal{E}_{H}(a_{-})=\langle a_{-}\,,\,
H_{\ell_{V}} a_{-}\rangle+\frac{G (\alpha_{1}+\alpha_{2}+2\alpha_{12})}{8}\int|a_{-}|^{4}$\,, in the limit
$\tau_{x}\to 0$\,.\\
Another possibility consists in considering the case when
$|\alpha_{2}-\alpha_{1}|+|\alpha_{12}-\alpha_{1}|$ is small as
$(\varepsilon,\tau_{x})\to (0,0)$\,. The energy
$\mathcal{E}_{\tau}(\psi)$\,, written as,
\begin{multline*}
\langle a_{-}\,,\,\hat{H}_{-}a_{-}\rangle
\\+\frac{G}{2}
\int
\left[\alpha_{1}+(\alpha_{2}-\alpha_{1})\sin^{4}(\frac{\theta}{2})+
\frac{(\alpha_{12}-\alpha_{1})}{2}\sin^{2}(\theta)\right]|a_{-}|^{4}~dxdy\,,
\end{multline*}
will converge to $\mathcal{E}_{H}(a_{-})=\langle a_{-}\,,\,
H_{\ell_{V}} a_{-}\rangle+\frac{G\alpha_{1}}{2}\int|a_{-}|^{4}$\,.
\item[2)] The existence of a minimizer for
$\mathcal{E}_{\varepsilon}$ and the variational argument showing that
$\|a_{+}\|_{L^{2}}^{2}=\mathcal{O}(\varepsilon^{2+4\delta})+\mathcal{E}_{\varepsilon,min}$ when
$\psi=\hat{U} \begin{pmatrix}
e^{i\frac{y}{\sqrt{\tau_{x}}}}a_{+}\\e^{-i\frac{y}{\sqrt{\tau_{x}}}}a_{-}
\end{pmatrix}$ is a ground state for $\mathcal{E}_{\varepsilon}$
will be essentially the same as in the isotropic case.
\item[3)] The analysis and the use of the Euler-Lagrange equation for
ground states of $\mathcal{E}_{\varepsilon}$, like in
Subsection~\ref{se.adELeq},
will certainly be more
delicate because it will be a system, and the vanishing of the
crossing terms have to be considered more carefully.
\end{description}
\subsection{Minimization for excited states}
\label{se.minexc}
One may consider like in \cite{DGJO} the question of minimizing the
energy, for states prepared according to $\psi_{+}$ local eigenvector
of the potential.
Two things have to be modified in order to adapt the previous
analysis:
\begin{description}
\item[1)] The space of states, on which the energy is
minimized has to be specified. The unitary transform $\hat{U}$
introduced in Theorem~\ref{th.BornOp} provides a simple way to
formulate this minimization problem: set
$\mathcal{F}_{+}=\Ran\hat{U}P_{+}$ with $P_{+}=
\begin{pmatrix}
1&0\\
0&0
\end{pmatrix}
$ and consider
$$
\inf_{\psi\in \mathcal{F}_{+}\,,\, \|\psi\|=1}\mathcal{E}_{\varepsilon}(\psi)\,.
$$
\item[2)] In order to get asymptotically as $\varepsilon\to 0$, the
same scalar minimization problems with $\mathcal{E}_{\tau}$ and
$\mathcal{E}_{H}$, the external potential $V_{\varepsilon,\tau}$ has
to be changed. It must be now
\begin{multline*}
V_{\varepsilon,\tau} (x,y)=
\frac{\varepsilon^{2+2\delta}}{\ell_{V}^{2}}v(\sqrt{\tau_{x}}x,\sqrt{\tau_{x}}y)
\\-
\sqrt{1+\tau_{x}x^{2}}-\varepsilon^{2+2\delta}
\left[
\frac{\tau_{x}^{2}}{(1+\tau_{x}x^{2})^{2}}+
\frac{1}{1+\tau_{x}x^{2}}
\right]\,.
\end{multline*}
\end{description}
The analysis of this problem is essentially the same as for the
complete minimization problem. It is even simpler because the unitary
$\hat{U}$ is directly introduced. A slightly different question is
about the minimization of the energy $\mathcal{E}_{\varepsilon}$
in the space $\mathcal{F}_{+}^{0}=\Ran
\hat{U}_{0}P_{+}$, but the accurate comparison between $\hat{U}$ and
$\hat{U}_{0}$ widely used through this article would lead to similar
results.\\
Possibly this extension can even be generalized to
higher rank matricial potentials with eigenvectors
$\psi_{1},\ldots, \psi_{N}$, for states modeled on any given
$\psi_{k}$\,.
\subsection{Time dynamics of adiabatically prepared states}
\label{se.timedyn}
Nonlinear adiabatic time evolution has been considered recently in
\cite{CaFe}.
Note that our problem is slightly different because, we are
considering a spatial adiabatic problem, but some techniques may be
related.\\
When $\psi_{0}=\hat{U}
\begin{pmatrix}
0\\alpha_{-,0}
\end{pmatrix}
$
(resp. $\psi_{0}=\hat{U}
\begin{pmatrix}
a_{+,0}\\0
\end{pmatrix}
$ ), the question is whether the solution $\psi(t)$ to
$$
i\partial_{t}\psi=H_{Lin}\psi+ G_{\varepsilon,\tau}|\psi|^{2}\psi\quad,\quad \psi(t=0)=\psi_{0}
$$
remains close to $\hat{U}
\begin{pmatrix}
0\\ a_{-}(t)
\end{pmatrix}
$
(resp. $
\begin{pmatrix}
a_{+}(t)\\0
\end{pmatrix}
$)
with
\begin{eqnarray*}
&&i\partial_{t}a_{-}=\varepsilon^{2+2\delta}\tau_{x}\hat{H}_{-}a_{-}+\frac{G\varepsilon^{2+2\delta}\tau_{x}}{2}|a_{-}|^{2}a_{-}\quad,\quad
a_{-}(t=0)=a_{-,0}\,,\\
\text{resp.}
&&
i\partial_{t}a_{+}=\varepsilon^{2+2\delta}\tau_{x}\hat{H}_{+}a_{+}+\frac{G\varepsilon^{2+2\delta}\tau_{x}}{2}|a_{+}|^{2}a_{+}\quad,\quad
a_{+}(t=0)=a_{+,0}\,.
\end{eqnarray*}
More precisely, the question is about the range of time where this
approximation is valid: what is the size of $T_{\varepsilon}$ w.r.t
$\varepsilon$ such that $\sup_{t\in [-T_{\varepsilon},T_{\varepsilon}]}\|\psi(t)-\hat{U}
\begin{pmatrix}
0\\alpha_{-}(t)
\end{pmatrix}
\|$ (resp. $\sup_{t\in [-T_{\varepsilon},T_{\varepsilon}]}\|\psi(t)-
\begin{pmatrix}
a_{+}(t)\\0
\end{pmatrix}
\|$) remains small.\\
Since the approximation of $\hat{U}^{*}H_{Lin}\hat{U}$ by
$\begin{pmatrix}
\hat{H}_{+}&0\\
0 &\hat{H_{-}}
\end{pmatrix}$ is good in the low frequency range, a natural
assumption will be that the initial data are supported in the low
frequency region
$$
a_{0,\pm}=\chi(\tau_{x}|\varepsilon D_{q}|^{2})a_{0,\pm}\,,
$$
for some compactly supported $\chi$\,. Then the question is whether the
norm of $(1-\tilde{\chi}(\varepsilon^{2}\tau_{x}D_{q}^{2}))\psi(t)$
remains small for $t\in [0,T_{\varepsilon}]$ for $\tilde{\chi}\in
\mathcal{C}^{\infty}_{0}(\field{R})$, $\chi\prec \tilde{\chi}$\,.
Then the two last parts of Section~\ref{se.minipb}, concerned with
the high frequency part and the effect of $\hat{U}$ on the
nonlinear term, have to be reconsidered.
\\
Note that the adiabatically prepared state with $\psi_{0}=\hat{U}
\begin{pmatrix}
a_{0,+}\\0
\end{pmatrix}
$
are probably not stable for
very long time,
$T_{\varepsilon}=\mathcal{O}(e^{\frac{c}{\varepsilon}})$, because the
characteristic set
$$
\mathcal{C}_{\lambda}=
\left\{(q,p)\in \field{R}^{4}, \quad
\det\left(\varepsilon^{2\delta}\tau_{x}p^{2}+V_{\varepsilon,\tau}(q)+M(q)-\lambda\right)
=0\right\}
$$
contains two components when $\lambda\geq \min E_{+}$, one
corresponding to the higher level of $M(q)$ with $|p|^{2}\leq
C(\lambda)$, and another one for the
lower level of $M(q)$ but with large $p$'s. This means that a tunnel
effect will occur between the two levels, so that adiabatically
prepared states, with energies close to $\lambda\geq \min E_{+}$, will
not remain in this state for (very) large times.
|
\section{Introduction}
Transition metal oxide clusters have attracted enormous attention
because their structural, electronic, and magnetic properties are
often quite different from those in their bulk phase
\cite{Nayak98,Reddy99,Harrwason00,Tono03,Pykavy04,Qu06,Uzunova08,WangQ08,Mowbray09,WangYB10}.
For example, small clusters like (ZnO)$_n$, (V$_2$O$_5$)$_n$,
(CrO$_3$)$_n$ form planar structures with very small sizes, and some
small clusters present novel magnetic properties
\cite{WangQ08,WangYB10,Vyboishchikov01,Li06}. For scandium (Sc),
continuous interests are shown for several reasons. Firstly, Sc is
the first transition metal in the periodic table, and thus is always
taken as a prototype to study the complex phenomena associated with
$d$ shell electrons \cite{Gonzales00}. Secondly, Sc oxide
nanostructures can be used as catalysts in some reactions like the
selective reduction of nitric oxides with methane
\cite{Gonzales00,Fokema98}. Thirdly, some Sc oxide clusters have
been recognized in the spectra of M-type stars
\cite{Gonzales00,Merrill62}. Finally, some Sc oxide clusters can be
steadily encapsulated into closed carbon cages of fullerenes to form
a novel nanostructure \cite{Stevenson08,Valencia09,Chaur09}.
The small Scandium oxide clusters can be prepared by laser ablation
of scandium metal in the presence of oxygen-saturated atmosphere
\cite{Chertihin97,Wu98,Kushto99,Zhao11}. With the presence of oxygen
or nitrogen oxide gases, scandium oxide clusters ranging from ScO to
Sc$_3$O$_6$ have already been generated and detected \cite{Zhao11}.
Many experimental studies have already been applied to study the
structural, energetic, vibrational, electronic, and magnetic
properties of scandium oxide clusters. For example, the
photoionization spectra of Sc$_n$O were measured and strongly
size-dependent ionization potentials were observed
\cite{Zhao11,Gutsev00}. For the ScO molecule, the electron-spin
resonance and optical spectroscopy in neon and argon matrices
revealed that its ground state is a doublet \cite{Zhao11,Weltner67},
and the molecular bond length and vibrational frequency have been
measured to be 1.668 \AA~ and 965 $cm^{-1}$ respectively
\cite{WangYB10,Huber79}. Accompanying with these experimental
results, lots of theoretical calculations, especially those based on
density functional theory (DFT), were also carried out to explore
the ground-state electronic structures of scandium oxide clusters.
However, most of those studies on Sc oxide clusters were mostly on
the standard DFT level, within the local density approximation (LDA)
or generalized gradient approximation (GGA). In this paper, we
systematically study the atomic and electronic structures of Sc
oxide clusters using the screened hybrid density functional theory,
which has proven to be able to significantly improve the description
of finite, molecular systems, such as atomization energies, bond
lengths and vibrational frequencies using standard DFT theories
\cite{Nieminen09}.
\section{Computational Method}
For extended systems, the ground-state electronic properties can be
obtained by solving the Kohn-Sham equation within DFT
\cite{Hohenberg64,Kohn65}, utilizing the standard approximations for
the exchange-correlation ($xc$) energy functional $E_{xc}$, i.e.,
LDA and GGA. One problem of the two most commonly applied
functionals is that they rely on the $xc$ energy per particle of the
uniform electron gas, and thus are expected to be useful only for
systems with slowly varying electron densities (local/semilocal
approximation) \cite{Hummer09}. Although the LDA and GGA functionals
have proven to be rather universally applicable in theoretical
materials science and achieve fairly good accuracy for ionization
energies of atoms, dissociation energies of molecules, and cohesive
energies, as well as bond lengths and geometries of molecules and
solids \cite{Hummer09,Gunnarsson76,Jones89,Kohn99}, this
unexpectedly good performance for the ground-state properties of
many materials is believed to be due to the partial error
cancelation in the exchange and correlation energies
\cite{Hummer09}. In order to improve the performance of DFT on a
theoretical basis, one approach is to add a portion of the non-local
Hartree-Fock (HF) exchange to the local/semilocal approximation for
exchange-correlation functional. In computational solid state
physics, a breakthrough was achieved by Heyd, Scuseria, and
Ernzerhof, who proposed the HSE03 functional defined by
\cite{Heyd03,Heyd04,Heyd06}:
\begin{equation}
E_{xc}^{\rm HSE03}=\frac{1}{4}E_{x}^{\rm HF}({\rm SR})+\frac{3}{4}E_{x}^{\rm PBE}({\rm SR})+E_{x}^{\rm PBE}({\rm LR})+E_{c}^{\rm PBE}.
\end{equation}
The HSE03 functional benefits from the range separation of the HF
exchange into a short-(SR) and long-range (LR) contribution and
replaces the latter by the corresponding DFT exchange part. The
HSE03 functional as well as its corrected form the HSE06 functional
\cite{Krukau06}, have been extensively applied to calculate
ground-state properties of solid and molecular systems, as well as
adsorption energies for molecules, and have proven to be better than
the LDA and GGA functionals in describing finite molecular systems.
The results presented in this work are obtained using the projector
augmented wave (PAW) method \cite{PAW} as implemented in the Vienna
ab initio simulation package (VASP) \cite{VASP}. For standard DFT
calculations the $xc$ energy is treated within the GGA using the
parameterization of Perdew, Burke, and Ernzerhof (PBE)
\cite{PBE1,PBE2} to compare with previously published calculations.
In the present HSE calculations, the HSE06 hybrid functional
\cite{Krukau06} is applied, where the range separation parameter is
set to 0.2 \AA$^{-1}$ for both the semilocal as well as the nonlocal
part of the exchange functional. From now on, this particular
functional will be referred to as HSE. A plane wave basis set with a
cutoff energy of 400 eV is adopted, and the scandium 3{\it
s}$^2$3{\it p}$^6$4{\it s}$^2$3{\it d}$^1$ and oxygen 2{\it
s}$^2$2{\it p}$^4$ electrons are treated as fully relaxed valence
electrons. A Fermi broadening \cite{Weinert1992} of 0.05 eV is
chosen to smear the occupation of the bands around the Fermi energy
(E$_f$) by a finite-$T$ Fermi function and extrapolating to $T=0$ K.
The supercell containing the scandium oxide clusters is chosen to be
without symmetry, and the cell size along each direction is larger
than 15 \AA. A quasi-Newton algorithm is used to relax the scandium
and oxygen ions for all scandium oxide clusters, with the force
convergence criteria of 0.01 eV/\AA~ in PBE calculations and 0.03
eV/\AA~ in HSE calculations.
\section{Results and Discussions}
\subsection{Geometrical structures}
Firstly, the ScO and two isomers of ScO$_2$ are shown in Fig.1 (a)
and (b) respectively. The optimized bond length of ScO is 1.677 \AA~
in PBE calculation, and 1.659 \AA~ in HSE calculation. The PBE
result is in good agreement with former GGA calculations
\cite{WangYB10}. The HSE06 result is near the ones obtained from the
SCF \cite{Dolg87} and hybrid B1LYP calculations
\cite{Qu06,Uzunova08}. Both GGA and HSE06 calculations agree well
with the experimental bond length (1.668\AA) \cite{Huber79}.
Moreover, The vibrational frequency of ScO is 1056 $cm^{-1}$ in HSE
calculation, which is near the experimental value of 965 $cm^{-1}$.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig1.eps}
\end{center}
\caption{(Color online). The low-energy structures of (a) ScO and
(b) ScO$_2$ clusters. Grey and red balls represent the scandium and
oxygen atoms, respectively. The bond lengths are in units of \AA.
The number in parenthesis is the relative energy (in eV) with
respect to the corresponding ground state. Note that the spin
multiplicity is also given in the parenthesis.}
\end{figure
For ScO$_2$, the obtained ground state is the obtuse triangle
structure, with the apex angle of 117.16$^{\circ}$. It is
interesting to note that these two Sc-O bonds are not equivalent and
the bond lengths are 1.687 and 1.928 \AA respectively. In PBE
calculations, the triangle is isosceles, with the apex angle of
126.19$^{\circ}$ and the ScO bond length of 1.781 \AA~, in agreement
with that have been found by using the DFT/PBE \cite{WangYB10} and
DFT/BPW91 methods \cite{Gutsev00}. It is clear that the HSE06
calculation reduces the symmetry of the ground state. Both GGA and
HSE06 demonstrates the doublet is the most stable electronic state
for the ground state of the ScO$_2$ cluster, in good agreement with
the experimental result. The next stable state of ScO$_2$ is the
O-Sc-O linear structure with 0.2~eV higher than the ground state.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig2.eps}
\end{center}
\caption{(Color online). The low-energy structures of (a) Sc$_2$O,
(b) Sc$_2$O$_2$, (c) Sc$_2$O$_3$, and (d) Sc$_2$O$_4$ clusters. Grey
and red balls represent the scandium and oxygen atoms, respectively.
The bond lengths are in units of \AA. The number in parenthesis is
the relative energy (in eV) with respect to the corresponding ground
state. Note that the spin multiplicity is also given in the
parenthesis.}
\end{figure
The low-lying isomers of Sc$_2$O$_m$ (m=1$\sim$4) are presented in
Fig.2 (a)-(d). The ground-state structure of Sc$_2$O cluster is also
an obtuse triangle. The apex angle is 109.51$^{\circ}$, and the two
Sc-O bonds are both 1.797 \AA~ long, indicating that the cluster has
a $C_{2v}$ point symmetry. The ground state of Sc$_2$O is a triplet,
which is in agreement with previous DFT studies \cite{WangYB10}. The
next stable structure of Sc$_2$O is linear, and has a triplet
electronic state. The Sc$_2$O$_2$ cluster adopts a distorted square
as its ground-state structure. The Sc-O bond lengths are 1.892,
1.698, 1.892 and 1.698 \AA~ respectively. Since the PBE calculation
prefers the rhombus with D$_{2h}$ symmetry, our result is in
accordance with the rule that the HSE06 functional tends to reduce
the point group symmetry of some clusters obtained from the PBE
calculations. The ground-state of Sc$_2$O$_2$ is singlet, and the
next stable state is the triplet state of the same structure. The
other two structures of Sc$_2$O$_2$ in Fig. 2(b) actually have much
higher electronic free energies. For Sc$_2$O$_3$, the ground-state
structure is the singlet trigonal bi-pyramid structure shown in Fig.
2(c). This structure has a $D_{3h}$ symmetry in PBE calculation,
which disappears in HSE calculation. Until now, our PBE
calculational results are in good agreements with the results
reported by Wang {\it et al.} \cite{WangYB10}. The main modification
of the HSE method, as we can see from the above description, is
revealing a lower symmetry for the ground-state structure. In the
previous work by Wang {\it et al.}, the ground-state structure of
Sc$_2$O$_4$ is the one with a molecular O$_2$ adsorbing at one
corner of the rhombus Sc$_2$O$_2$ cluster, which is however, proven
to be the next stable structure for Sc$_2$O$_4$. The more stable
structure we revealed for Sc$_2$O$_4$ is the adsorbing structure
with an O$_2$ molecule adsorbing on top of the Sc-Sc bond, as shown
in Fig. 2(d). The ground and next stable states of Sc$_2$O$_4$ are
both singlet.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig3.eps}
\end{center}
\caption{(Color online). The low-energy structures of (a) Sc$_3$O,
(b) Sc$_3$O$_2$, (c) Sc$_3$O$_3$, and (d) Sc$_3$O$_4$ clusters. Grey
and red balls represent the scandium and oxygen atoms, respectively.
The bond lengths are in units of \AA. The number in parenthesis is
the relative energy (in eV) with respect to the corresponding ground
state. Note that the spin multiplicity is also given in the
parenthesis.}
\end{figure
The Sc$_3$O cluster has been reported to have a singlet ground
state, with the structure of an oxygen atom adsorbing on top of an
equilateral scandium triangle \cite{WangJL09}. Our result of Sc$_3$O
is the same with previous, in both PBE and HSE calculations. The
next stable structure of Sc$_3$O is the one with an oxygen atom
adsorbing at the bottom of an isosceles scandium triangle, having a
singlet electronic state and a free electronic energy of 0.16 eV
higher than that of the ground-state of Sc$_3$O. Except for Sc$_3$O,
other Sc$_3$O$_m$ clusters have seldom been studied before. Only a
systematic study on the Sc$_3$O$_6$ cluster is reported very
recently \cite{Zhao11}. The ground-state of Sc$_3$O$_2$ is the
doublet state of the adsorbing structure of a scandium atom on the
Sc$_2$O$_2$ cluster. We also notice that there are several structure
for the Sc$_3$O$_2$ cluster having close free electronic energies,
which are thus shown in Fig. 3(b). They are the pentagon, bi-pyramid
and bi-triangle structures respectively. The Sc$_3$O$_3$ cluster
adopts a hexagon structure with a fourfold spin multiplicity as its
ground state. The smallest O-Sc-O angle is 101.59$^{\circ}$ in PBE
calculation, and 102.78$^{\circ}$ in HSE calculation. We also listed
several other structures of Sc$_3$O$_3$ in Fig. 3(c), especially the
next stable doublet state of the folded rectangular structure, which
is only 0.04 eV higher in free electronic energy. For Sc$_3$O$_4$,
the ground-state structure can be seen as the adsorption of an
oxygen atom at the center of the hexagonal Sc$_3$O$_3$ cluster, or
the adsorption of three oxygen atoms at the three bottom sides of
the pyramid Sc$_3$O cluster. The next stable structure of
Sc$_3$O$_4$, which is 0.42 eV higher in free electronic energy, can
be seen as replacing one oxygen atom in the hexagonal Sc$_3$O$_3$
cluster with two oxygen atoms.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig4.eps}
\end{center}
\caption{(Color online). The low-energy structures of (a)
Sc$_3$O$_5$ and (b) Sc$_3$O$_6$ clusters. Grey and red balls
represent the scandium and oxygen atoms, respectively. The bond
lengths are in units of \AA. The number in parenthesis is the
relative energy (in eV) with respect to the corresponding ground
state. Note that the spin multiplicity is also given in the
parenthesis.}
\end{figure
As shown in Fig. 4(a), the ground-state structure for Sc$_3$O$_5$
can be seen as adding an oxygen atom on top of the scandium atom in
the ground-state structure of Sc$_3$O$_4$. The Sc-O bond lengths
become a little smaller than that in Sc$_3$O$_4$. The next stable
structure of Sc$_3$O$_5$ is the adsorption of 5 oxygen atoms on a
Sc$_3$ linear chain, whose free electronic energy is however, much
larger than the ground-state. The 3 lowest-energy structures of
Sc$_3$O$_6$ are the adsorption structures of 6 oxygen atoms on the
Sc$_3$ triangle. In the ground state, the oxygen atoms are separated
from each other, while in the next and third stable structure, two
of the six oxygen atoms bond together. We also find that the
structures contain a Sc$_3$ chain always have much larger free
electronic energies.
\subsection{Important features from the structural studies}
Based on the above systematic study of Sc$_n$O$_m$ clusters, we can
see some features for both the oxidation pattern of scandium, and
the influences of HSE calculation.
For the oxidation of scandium, we conclude two important features.
The first one is that the Sc$_n$O$_m$ cluster can be seen as adding
$m$ oxygen atoms to a Sc$_n$ cluster, with the Sc$_n$ cluster to be
scandium dimer for $n$=2, and scandium triangle for $n$=3. The other
feature is that the oxygen atoms in the ground-states of Sc$_n$O$_m$
clusters are all bondless with each other, only one O-O bond forms
in Sc$_2$O$_4$. The fact that oxygen atoms do not bond with each
other indicates that scandium oxide clusters are different from lead
oxide clusters, in which oxygen atoms can form O$_2$ or O$_3$ units
\cite{Liu07}. The structural features of Sc$_n$O$_m$ clusters tell
us that they can be built by adding $m$ oxygen atoms separately to a
Sc$_n$ cluster.
As for the improvements of the HSE method, the most important one is
that the symmetry of the cluster is often lower in HSE calculation
than in PBE calculation. The lengths of Sc-O bonds always differ
from each other in a cluster after geometry optimization using the
HSE method. The second feature concluded from the above geometrical
studies is that the HSE method does not change the relative
stability between different structures. It means that geometry
optimization using the PBE type functional will not yield wrong
ground-state structures for scandium oxide clusters. At last,
although the electronic structures in PBE and HSE calculations are
the same for most Sc$_n$O$_m$ clusters, we do see differences in the
ground-state electronic structures of Sc$_3$O$_2$ and Sc$_3$O$_3$.
In PBE calculations, the magnetic moment is 3 $\mu_B$ for
Sc$_3$O$_2$, and 1 $\mu_B$ for Sc$_3$O$_3$. And corresponding the
magnetic moments obtained from HSE calculations are 1 $\mu_B$ for
Sc$_3$O$_2$, and 3 $\mu_B$ for Sc$_3$O$_3$.
\subsection{Fragmentation channels and dissociation energies}
The stability of scandium oxide clusters with different sizes and
stoichiometries is required to illustrate the growth pattern of
various nanostructures and to understand even the oxidation behavior
of the pristine scandium clusters. On the other hand, the study of
the stability is helpful for finding the candidates of the building
block of the cluster-assembled materials. Therefore, in the
following, we evaluate the fragmentation energy of Sc$_n$O$_m$
clusters.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig5.eps}
\end{center}
\caption{(Color online). Dissociation energies of the Sc$_n$O$_m$
clusters as a function of the oxygen atom number $m$. The
corresponding fragmentation channels are listed in Table I.}
\end{figure
\begin{table}[ptb]
\caption{The most favorable fragmentation channels and dissociation
energies ($\Delta E$, in units of eV) of the Sc$_n$O$_m$ ($1\leq
n\leq 3,~1\leq m\leq 2n$) clusters. All results are obtained by
employing the HSE method.}
\begin{tabular}
[c] { l l l} \hline \hline
cluster ~~~~& Fragmentation channels ~~~~& $\Delta E$ \\
\hline
ScO & Sc+O & +4.29 \\
ScO$_2$ & ScO+O & +1.41 \\
Sc$_2$O & Sc+ScO & +2.57 \\
Sc$_2$O$_2$ & ScO+ScO & +4.45 \\
Sc$_2$O$_3$ & Sc$_2$O$_2$+O & +3.79 \\
Sc$_2$O$_4$ & Sc$_2$O$_3$+O & +0.12 \\
Sc$_3$O & Sc+Sc$_2$O & +1.78 \\
Sc$_3$O$_2$ & Sc+Sc$_2$O$_2$ & +1.80 \\
Sc$_3$O$_3$ & ScO+Sc$_2$O$_2$ & +3.60 \\
Sc$_3$O$_4$ & ScO+Sc$_2$O$_3$ & +5.60 \\
Sc$_3$O$_5$ & Sc$_3$O$_4$+O & +2.03 \\
Sc$_3$O$_6$ & Sc$_3$O$_5$+O & +1.88 \\
\hline \hline
\end{tabular}\label{Ead1}
\end{table}
When a cluster $A$ is dissociated into $B$ and $C$ fragments (i.e.,
$A\rightarrow B+C$), the fragmentation energy is defined as $\Delta
E =E_B+E_C-E_A$, where the $E_B$, $E_C$, and $E_A$ are the free
electronic energies of clusters $B$, $C$, and $A$, respectively. In
addition, the half of the total energy of an oxygen molecule is
adopted as the reference energy for atomic oxygen. Note that a
fragmentation process is exothermic (endothermic) if the associated
fragmentation energy is negative (positive). The dissociation energy
$\Delta E$, defined as the fragmentation energy of the most
favorable channel, is calculated and presented in Table I. In
general, the cluster with large positive $\Delta E$ has great
stability, and that with small positive or even negative $\Delta E$
is not stable and tends to dissociate. Moreover, the frequently
observed fragmentation products are believed to be stable
\cite{Liu07,Lu03}. Figure 5 shows the $\Delta E$ as a function of
the oxygen atom numbers for for all clusters, revealing the
underlying relationship between stability and stoichiometry of these
clusters. The essential features can be discussed as follows.
Firstly, for the dissociation of Sc-rich clusters, the Sc atom is
always one of the fragments. To make a further validation, we also
study possible fragmentation ways for the Sc$_4$O cluster, using the
HSE calculational method. It is found that the most energetically
favored fragmentation channel is Sc$_4$O$\rightarrow$Sc+Sc$_3$O, in
which single Sc atom is also a fragment. Secondly, we do not see
oxygen molecules in the fragments of Sc$_n$O$_m$ clusters, the
O-rich clusters favor the fragmentation channels that have a single
oxygen atom as a product. Lastly, the dissociation energies of ScO,
Sc$_2$O$_2$, Sc$_2$O$_3$, Sc$_3$O$_3$, and Sc$_3$O$_4$ clusters are
larger than 3.60 eV. This indicates that the monoxide-like and
sesquioxide-like scandium oxide clusters, and oxide clusters between
them are remarkably stable. As a result, these clusters are
frequently observed in the fragmentation channels shown in Table I.
We also notice that the Sc$_3$O$_4$ cluster has an enormous
dissociation energy, and is more stable than the monoxide-like
Sc$_3$O$_3$ cluster. And at another side, the monoxide-like
Sc$_2$O$_2$ cluster, with a larger dissociation energy, is more
stable than the the sesquioxide-like Sc$_2$O$_3$ cluster.
\subsection{Electronic energy levels}
The electronic structures of transition metal oxide clusters are
important characters for their experimental detections and chemical
applications. Especially, their magnetic properties are very
important to reflect the behaviors of the $d$-shell electrons. We
thus calculate the ground-state electronic energy levels of
Sc$_n$O$_m$ clusters, using the HSE method. Figure 6 shows the
obtained energy levels for different spin polarized electrons, in
which spin-up and spin-down represent for the majority and minority
spin states respectively. We find that the Sc$_n$O$_m$ clusters with
odd-number Sc atoms are all magnetic. For the Sc$_2$O$_m$ clusters,
while the Sc$_2$O is magnetic, the other Sc$_2$O$_2$, Sc$_2$O$_3$,
and Sc$_2$O$_4$ clusters are nonmagnetic. The energy gap between the
highest occupied molecular orbital (HOMO) and the lowest unoccupied
molecular orbital (LUMO) is 1.82, 3.36, and 3.20 eV for the
Sc$_2$O$_2$, Sc$_2$O$_3$, and Sc$_2$O$_4$ cluster respectively. The
large band gaps indicate that the Sc$_2$O$_2$, Sc$_2$O$_3$ and
Sc$_2$O$_4$ clusters are chemically very stable. In contrast,
although the Sc$_3$O$_4$ cluster is energetically very stable, its
energy gap of spin-up electrons is as small as 0.49 eV, indicating
that it is chemically active, and ready to interact with other
particles or materials.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig6.eps}
\end{center}
\caption{(Color online). The electronic energy levels for (a) ScO,
(b) ScO$_2$, (c) Sc$_2$O, (d) Sc$_2$O$_2$, (e) Sc$_2$O$_3$, (f)
Sc$_2$O$_4$, (g) Sc$_3$O, (h) Sc$_3$O$_2$, (i) Sc$_3$O$_3$, (j)
Sc$_3$O$_4$, (k) Sc$_3$O$_5$, and (l) Sc$_3$O$_6$ clusters, obtained
from HSE calculations. Spin-up and spin-down electronic states are
denoted by black and red lines, respectively. The Fermi levels are
denoted by the blue dotted lines.}
\end{figure
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.3\textwidth]{fig7a.eps}
\includegraphics[width=0.3\textwidth]{fig7b.eps}
\includegraphics[width=0.3\textwidth]{fig7c.eps}
\includegraphics[width=0.3\textwidth]{fig7d.eps}
\includegraphics[width=0.3\textwidth]{fig7e.eps}
\includegraphics[width=0.3\textwidth]{fig7f.eps}
\includegraphics[width=0.3\textwidth]{fig7g.eps}
\includegraphics[width=0.3\textwidth]{fig7h.eps}
\includegraphics[width=0.3\textwidth]{fig7i.eps}
\includegraphics[width=0.3\textwidth]{fig7j.eps}
\includegraphics[width=0.3\textwidth]{fig7k.eps}
\includegraphics[width=0.3\textwidth]{fig7l.eps}
\end{center}
\caption{(Color online). The spin-resolved projected density of
states for the (a) ScO, (b) ScO$_2$, (c) Sc$_2$O, (d) Sc$_2$O$_2$,
(e) Sc$_2$O$_3$, (f) Sc$_2$O$_4$, (g) Sc$_3$O, (h) Sc$_3$O$_2$, (i)
Sc$_3$O$_3$, (j) Sc$_3$O$_4$, (k) Sc$_3$O$_5$, and (l) Sc$_3$O$_6$
clusters. The Fermi energies are all set to be zero. The electronic
states of O and Sc are shown in dotted and solid line respectively,
while $s$, $p$ and $d$ electronic states are shown as black, blue
and red lines respectively.}
\end{figure
To systematically investigate the electronic structures of
Sc$_n$O$_m$ clusters, we then calculate the projected density of
states (PDOS) for the Sc$_n$O$_m$ clusters, which are spin-resolved.
Figures 7(a)-7(l) list the obtained $s$-, $p$-, and $d$-PDOS of
oxygen and scandium atoms, for the Sc$_n$O$_m$ clusters
respectively. We can clearly see that for all the studied
Sc$_n$O$_m$ clusters, there are strong hybridizations between oxygen
2$p$ and scandium 3$d$ electronic states, and they contribute most
of the electronic states around the Fermi energies. For Sc-rich
clusters like Sc$_2$O, Sc$_3$O, and Sc$_3$O$_2$, the HOMO and LUMO
are both composed of Sc-3$d$ electrons. For the three nonmagnetic
Sc$_2$O$_2$, Sc$_2$O$_3$, and Sc$_2$O$_4$ clusters, the HOMO and
LUMO are contributed by O-2$p$ electrons and Sc-3$d$ electrons,
respectively. And for O-rich clusters like ScO$_2$, Sc$_3$O$_5$, and
Sc$_3$O$_6$, the HOMO and LUMO are both 2$p$ electronic states of
oxygen. The electronic structure of ScO is, however, different from
the other clusters, for its HOMO of spin-up electrons and LUMO of
spin-down electrons are contributed by 4$s$ electronic states of
scandium.
\begin{figure}[tbp]
\begin{center}
\includegraphics[width=0.8\linewidth]{fig8.eps}
\end{center}
\caption{(Color online). The spin-resolved projected density of
states for the Sc$_3$O$_2$ cluster in (a) PBE and (b) HSE
calculations, and the Sc$_3$O$_3$ cluster in (c) PBE and (d) HSE
calculations. The Fermi energies are all set to be zero. The
electronic states of O and Sc are shown in dotted and solid line
respectively, while $s$, $p$ and $d$ electronic states are shown as
black, blue and red lines respectively.}
\end{figure
Since the obtained ground-state spin configurations are different
for the Sc$_3$O$_2$ and Sc$_3$O$_3$ clusters in PBE and HSE
calculations, we draw their PDOS together to compare the differences
in electronic-state descriptions of PBE and HSE methods, which are
shown in Figs. 8(a)-8(d). In the $p$-$d$ hybridization area from
-7.0 eV to -3.0 eV below the Fermi energies, we can see that the
hybridized peaks are more separate in HSE calculations for both the
Sc$_3$O$_2$ and Sc$_3$O$_3$ clusters. The electronic states around
the Fermi energy are both Sc-3$d$ states for the two chosen
clusters. One can see from Fig. 8 that there are always more
localized Sc-3$d$ peaks around the Fermi energy in HSE calculations
than in PBE calculations. It means that the HSE method can lead to
more localized descriptions for Sc-3$d$ electronic states in
Sc$_n$O$_m$ clusters. Considering that the PBE and other standard
GGA methods rely on the $xc$ energy of the uniform electron gas, and
thus are expected to be useful only for systems with slowly varying
electron densities, we think that the HSE descriptions on the
Sc-3$d$ states are more reasonable.
\section{Conclusions}
In this work, first-principles calculations with the HSE method have
been performed to study the geometries, stabilities and electronic
structures of small Sc$_n$O$_m$ clusters ($n$=1-3,$m$=1-2$n$). Based
on an extensive search, it is found that the lowest-energy
structures of all these clusters can be obtained by the sequential
oxidation of small "core" scandium clusters, with the adsorbing
oxygen atoms separating from each other.
The fragmentation analysis reveals that the ScO, Sc$_2$O$_2$,
Sc$_2$O$_3$, Sc$_3$O$_3$, and Sc$_3$O$_4$ clusters have great
stability. This suggests that these clusters might be used as
candidates of the building block of cluster-assembled materials. The
fragmentation of Sc-rich clusters is found to include a scandium
atom as a product, while the fragmentation of O-rich clusters is
found to include an oxygen atom as a product. Besides, the above
four extremely stable clusters can also be frequently seen in the
fragmentation products of other Sc$_n$O$_m$ clusters.
Through electronic structure calculations and wavefunction analysis,
we reveal that most Sc oxide clusters have magnetic ground state
except Sc$_2$O$_2$, Sc$_2$O$_3$, and Sc$_2$O$_4$ clusters. The HOMO
and LUMO of the three nonmagnetic clusters are composed of O-2$p$
and Sc-3$d$ electronic states, respectively. For the Sc-rich
clusters, the HOMO and LUMO are contributed both by Sc-3$d$
electrons, while for O-rich clusters, the HOMO and LUMO are
contributed by O-2$p$ electrons.
In comparison with standard PBE calculations, the HSE method is
superior because it can correct the wrong symmetries and electronic
configurations in PBE results of some clusters.
\begin{acknowledgments}
This work was supported by the NSFC under Grants No. 90921003 and
No. 10904004.
\end{acknowledgments}
|
\section{Introduction}
Over 550 exoplanets have been discovered to date. Many of the planets orbit solar-mass (0.7$-$1.5 $M_{\odot}$) stars, and they have revealed properties that are now used to constrain planet-formation models (e.g., \cite{Ida2004}; \cite{Butler2006}; \cite{Udry2007}). In contrast, only about 60 and 25 planets have been detected around evolved G-K (sub)giants (1.5$-$5 $M_{\odot}$) and K-M dwarfs ($<$0.7 $M_{\odot}$), respectively (e.g., \cite{Sato2008}; \cite{Johnson2011}; \cite{Johnson2007a}). Accordingly, the properties of the planetary systems orbiting such stars are less clarified yet than those for solar-mass stars. Planetary formation depends on the properties of protoplanetary disks, which should be affected by properties of the host star, such as stellar metallicity, radiation output, and disk diffusion times (e.g., \cite{Kornet2006}; \cite{Kennedy2008}). Observational features of planetary systems over a wide range of host star masses need to be clarified by current and future surveys of various masses stars in order to understand planetary formation in general.
More than 20 years ago, initial theoretical ideas of planetary formation for systems over a wide range of stellar masses were presented in terms of planet formation in protoplanetary disks with different properties (\cite{Nakano1988a}; \cite{Nakano1988b}). In the last two decades, improvements in planet formation modeling have made it possible to compare theoretical models directly with observed properties of planetary systems around stars with various masses (e.g., \cite{Ida2005}; \cite{Burkert2007}). For example, \citet{Kennedy2008} predicted that the peak occurrence rate of giant planets occurs for stars with masses of around 3 $M_{\odot}$, based on a core accretion scenario which includes the movements of snow lines under the evolution of central stars. Moreover, \citet{Currie2009} suggested that "the planet desert", i.e., a dearth of planets with semi-major axes of $<$0.6 AU orbiting $>$1.5 $M_{\odot}$ stars, may be reproduced by the effects of Type-II migration, considering the dependence of diffusion time of the protoplanetary disk on stellar mass. Clarifying the relationship between stellar mass and planetary system will provide valuable insights into planet formation models.
For intermediate-mass stars on the main sequence, precise Doppler surveys are difficult because of their large intrinsic radial velocity variations and smooth spectra with few absorption lines, caused by high surface activity, high surface temperature and/or high rotational velocity \citep{Lagrange2009}. In contrast, evolved intermediate-mass (sub)giant stars are suitable targets for precise Doppler surveys because these stars have low surface activity and their spectra exhibit many sharp absorption lines. Thus, to date, spectroscopy-based planet searches targeting intermediate-mass stars have been carried out through precise Doppler surveys of evolved stars. Although the number of substellar companions found orbiting such stars is still insufficient, some characteristic planetary system properties across a wide range of host star masses have begun to emerge. For example, the masses of planets and their host stars show correlation: more massive substellar companions tend to exist around more massive stars (e.g., \cite{Lovis2007}). This correlation suggests that the mass range of the brown dwarf desert depends on host-star's mass, and that planets may be deficient around 2.4$-$4 $M_{\odot}$ stars \citep{Omiya2009}. Also, the planet occurrence rate depends on host-star's mass: the giant planet frequency for higher-mass giant stars is higher than that for lower-mass stars (\cite{Lovis2007}; \cite{Johnson2007a}). The fraction of giant planets increases with increasing stellar mass up to 2 $M_{\odot}$ \citep{Johnson2010a}. Moreover, the orbital semi-major axes of planetary systems also seem to be correlated to host-star's properties. Semi-major axes of most planets orbiting intermediate-mass (sub)giant stars are larger than 0.6 AU\footnote[1]{A planet with a semi-major axis of 0.081 AU was found orbiting an intermediate-mass subgiant star HD 102956 with a mass of 1.68 $M_{\odot}$ \citep{Johnson2010b}.}, while those orbiting solar-type stars are larger than 0.02 AU (\cite{Johnson2007b}; \cite{Sato2008}; \cite{Wright2009}, \cite{Bowler2010}). Even considering the effect of engulfment of inner-orbit planets by host stars, which have experienced rapid expansion in the red giant branch (RGB) phase (\cite{Sato2008}, \cite{Villaver2009}), the observed properties of substellar systems orbiting intermediate-mass (sub)giant stars seem to be different from those orbiting solar-type stars (see also \cite{Bowler2010}).
In 2005, we started a Doppler spectroscopy-based survey of evolved GK-type giants in a framework of a Korean$-$Japanese planet search program \citep{Omiya2009}. The survey program is an extension to the ongoing Okayama Astrophysical Observatory (OAO) planet search program \citep{Sato2005}, and aims to clarify the properties of their associated planetary systems in collaboration with an East-Asian Planet Search Network (EAPS-Net; \cite{Izumiura2005}). About 190 sample stars of the survey were selected from the $Hipparcos$ catalog based on the same criteria as those for OAO planet search program, except visual magnitude (6.2 $<$ $V$ $<$ 6.5). The radial velocity variability of each sample star is monitored using either the 1.8-m telescope at Bohyunsan Optical Astronomy Observatory (BOAO, Korea) or the 1.88-m telescope at OAO (Japan). If a sample star exhibits large variations in radial velocity, follow-up observations of the star are performed using both telescopes.
In this paper, we report the discovery of a planetary companion orbiting the intermediate-mass giant HD 100655. This is the first planet discovered by this Korean-Japanese planet search program. In section 2, we describe our observations and radial velocity measurements from BOAO and OAO data. The properties of the host star and the radial velocity variability are reported in sections 3 and 4, respectively. We discuss possible causes of the radial velocity variation in section 5. In section 6, we consider the implications of this discovery for the current picture of planetary companions around intermediate-mass giant stars.
\section{Observations and Analyses}
\subsection{BOES Observations and Analysis}
Radial velocity observations at BOAO were carried out with the 1.8-m telescope and the BOAO Echelle Spectrograph (BOES; \cite{Kim2007}), a fiber-fed high resolution echelle spectrograph. We placed an iodine (I$_{2}$) cell in the optical path in front of the fiber entrance of the spectrograph \citep{Kim2002} for precise wavelength calibration and used a 200-$\mu$m fiber, obtaining a wavelength resolution $R$ = $\lambda$/$\Delta{\lambda}$ $\sim$ 51,000. The spectra covered a wavelength region from 3500 $\mathrm{\AA}$ to 10,500 $\mathrm{\AA}$. Echelle data reduction was performed using the IRAF\footnote[2]{IRAF are 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, USA.} software package in the standard manner. We used a wavelength region of 5000$-$5900 $\mathrm{\AA}$ which is covered by many I$_{2}$ absorption lines, for radial velocity measurements. We also made use of Ca II H line at around 3970 $\mathrm{\AA}$ as chromospheric activity diagnostics. Radial velocity analysis was performed using the spectral modeling technique described in \citet{Sato2002}, which was based on the method of \citet{Butler1996} and was adapted to BOES data analysis \citep{Omiya2009}. We employed the extraction method described in \citet{Sato2002} to prepare a stellar template spectrum from stellar spectra taken through the I$_{2}$ cell (star+I$_{2}$ spectra). The technique allowed us to achieve a long-term Doppler precision of 14 m s$^{-1}$ over 4.5 years.
\subsection{HIDES Observations and Analysis}
Radial velocity observations at OAO were carried out with the 1.88-m telescope and HIgh Dispersion Echelle Spectrograph (HIDES; \cite{Izumiura1999}) attached to the coud\'e focus of the telescope. We used an I$_{2}$ cell placed in the optical path in front of the slit of the spectrograph \citep{Kambe2002} as a precise wavelength calibrator. We always set the slit width to 200 $\mu$m (0.76"), providing a spectral resolution of 63,000. Until November 2007, we had taken star+I$_{2}$ spectra with a wavelength region of 5000$-$6200-$\mathrm{\AA}$. Since the HIDES CCD system was upgraded to a three-CCD mosaic in December 2007, we have obtained spectra from 3750 $\mathrm{\AA}$ to 7550 $\mathrm{\AA}$. The wavelength region of 5000$-$5900 $\mathrm{\AA}$ of the star+I$_{2}$ spectra are used for radial velocity measurements. The full range of stellar spectra taken without the I$_{2}$ cell are used for abundance analysis. Echelle data reduction was performed using the IRAF software package in the standard manner. Stellar radial velocities were derived from the star+I$_{2}$ spectra using the spectral modeling techniques detailed in \citet{Sato2002}, giving a Doppler precision of less than 8 m s$^{-1}$ over 4.5 years.
\section{Stellar Parameters of HD 100655}
HD 100655 (HR 4459, HIP 56508, BD+21 2331) is 122.3 $\pm$ 7.5 pc from the Sun according to the $Hipparcos$ parallax of $\pi$ = 8.18 $\pm$ 0.50 mas \citep{vanLeeuwen2007}. The star is classified as a G9III giant star with $V$ = 6.45 and $B-V$ = 1.010 $\pm$ 0.015 \citep{ESA1997}. We corrected the observed color index by an extinction value of $E(B-V)$ = 0.0163 $\pm$ 0.0016. The value was calculated from the galactic extinction of $E(B-V)_{S}$ = 0.0273 $\pm$ 0.0015 to the direction of the star obtained from the \citet{Schlegel1998} dust maps using the relation $E(B-V)$ = $E(B-V)_{S}$[1$-$exp($-|D$sin$b|$/125)], where $D$ and $b$ are the distance from sun and the galactic latitude, respectively. We derived an effective temperature of the star of $T_\mathrm{eff}$ = 4861 $\pm$ 110 K using the $(B-V)-T_{\mathrm{eff}}$ calibration of Alonso et al. (\yearcite{Alonso1999}, \yearcite{Alonso2001}). A luminosity of $L$ = 43 $\pm$ 5 $L_{\odot}$ was obtained from the absolute magnitude $M_{V}$ = 0.96 $\pm$ 0.13 and the bolometric correction $B.C.$ = $-$0.31 $\pm$ 0.04 based on the calibration of \citet{Alonso1999}. A stellar mass of $M$ = 2.4$^{+0.2}_{-0.4}$ $M_{\odot}$ was estimated by interpolating the evolutionary tracks of \citet{Girardi2000} with the estimated $T_{\mathrm{eff}}$ and $L$ (see figure \ref{fig1}). We determined the surface gravity to be log $g$ = 2.89 $\pm$ 0.10 and the stellar radius $R$ = 9.3$^{+1.3}_{-1.1}$ $R_{\odot}$ from $M$, $L$, and $T_{\mathrm{eff}}$. The microturbulent velocity $V_{t}$ = 1.36 $\pm$ 0.03 km s$^{-1}$ and the [Fe/H] of 0.15 $\pm$ 0.12 were derived from abundance analysis of a model atmosphere \citep{Kurucz1993} using the equivalent widths of Fe I and Fe II lines measured from an I$_{2}$-free spectrum of HD 100655. We adopted gf-values of Fe I and Fe II lines from \citet{Takeda2005}. \citet{deMedeiros1999} found the stellar rotational velocity, $v\mathrm{sin} i_{s}$, to be 1.6 $\pm$ 1.0 km s$^{-1}$. This value is comparable to the rotational velocities of typical late G-type giants. The stellar parameters are summarized in table \ref{tab1}.
\section{Orbital Solution}
A large radial velocity variation in the star HD 100655 was found in the early BOAO survey and we made intensive follow-up observations of the star at BOAO and OAO. For 4.5 years from the beginning of the survey, we collected 13 BOAO data points having a typical signal-to-noise ratio (S/N) of 170 pixel$^{-1}$ with an exposure time of 900$-$1200 s, and 32 OAO data points having a typical S/N of 120 pixel$^{-1}$ with an exposure time of 1200$-$1800 s. The observed radial velocities of HD 100655 are shown in figure \ref{fig2} and listed in table \ref{tab2}, together with the observation dates (JD) and estimated uncertainties. A dominant peak in the Lomb-Scargle periodogram \citep{Scargle1982} of the radial velocity variation exists at a period of 157.78 d (a frequency of 0.006338 c d$^{-1}$) (see figure \ref{fig3}). To check the significance of this periodicity, we estimated a False Alarm Probability ($FAP$) using the bootstrap randomization method. We produced 10$^{5}$ fake data sets by randomly mixing the observed radial velocities with a fixed observation date, and applied the Lomb-Scargle periodogram analysis to them. Only one fake data set showed a periodogram power higher than the observed one. Thus, the $FAP$ of the period is 10$^{-5}$. A best-fit Keplerian orbit derived from both the BOAO and OAO velocity data by a least-squares fit has a period $P$ $=$ 157.57 d, a velocity semi-amplitude $K_{1}$ $=$ 35.2 m s$^{-1}$, and an eccentricity $e$ $=$ 0.085. The best-fit curve is shown in figure \ref{fig2} as a solid line overlaid on the observed velocities. We applied an offset of $\Delta$RV = $-$28.1 m s$^{-1}$ to the BOAO velocity data, estimated concurrently with the orbital fit to a Keplerian model. The offset was required because of difference of velocity zero points between BOES and HIDES data originated from using different stellar templates for each data. The rms of the residuals to the best-fit are 14.9 m s$^{-1}$ for BOAO data, 9.2 m s$^{-1}$ for OAO data, and 11.2 m s$^{-1}$ for combined data sets. In the residuals we could not find any significant periodic variation due to additional companions. The best-fit orbital parameters and their uncertainties are listed in table \ref{tab3}. The uncertainties were estimated using a bootstrap Monte Carlo approach by creating 1000 fake data sets. Adopting a stellar mass $M$ = 2.4$^{+0.2}_{-0.4}$ $M_{\odot}$ for HD 100655, we obtained a semi-major axis $a$ = 0.76$^{+0.02}_{-0.04}$ AU and a minimum mass $M_{\mathrm{2}} \mathrm{sin} i_{p}$ = 1.7$^{+0.1}_{-0.2}$ $M_{\mathrm{J}}$ for the planetary companion.
\section{Cause of the Radial Velocity Variation}
To examine causes of the apparent radial velocity variation other than orbital motion, we checked the Ca II H line and the $Hipparcos$ photometric variation, and performed spectral-line shape analyses using a technique described in \citet{Sato2007} as follows. In the analyses, we investigated the cause of the velocity difference between spectra observed at top and bottom velocity phase.
Figure \ref{fig4} shows the spectrum around the Ca II H line of HD 100655. We note a lack of significant emission in the Ca II H line core of HD 100655, which suggests chromospheric inactivity for the star. Moreover, $Hipparcos$ photometry demonstrates the photometric stability of HD 100655 down to $\sigma$ $\sim$ 0.008 mag. based on the 55 observations for the star over a period of 1000 d. Figure \ref{fig5} displays a periodogram of the $Hipparcos$ photometry. We note a weak peak around the period of the radial velocity variation. To check the significance of the peak, we estimated $FAP$ using the bootstrap method as well as the method described in section 4. We produced 10$^{5}$ fake data sets, and applied the Lomb-Scargle periodogram analysis to them. A total of 7425 fake datasets showed a peak around the period of the radial velocity variation higher than the peak on the observed data set, which means $FAP$ of the peak is about 7.4 \%. Thus the peak is not considered to be significant. Although we have not completely disproved the possibility that the radial velocity variation is due to rotational modulation, these photometric results suggest that the main cause of the observed radial velocity variation is not rotational modulation of stellar spots.
For spectral-line shape analysis, we extracted two high-resolution stellar templates from star(HD 100655)+I$_{2}$ spectra obtained at OAO, using the method described in \citet{Sato2002}. One template was constructed from four spectra with observed radial velocities of the peak phase ($\sim$32 m s$^{-1}$), and the other from four spectra of the valley phase ($-$44 m s$^{-1}$ to $-$34 m s$^{-1}$). Cross-correlation profiles of the two templates were provided for 75 spectral segments (4-\AA \ to 5-\AA \ width each) that did not include severely blended lines or broad lines. We obtained a bisector for the cross-correlation profile of each segment and calculated three quantities from velocities at three flux levels (25\%, 50\%, and 75\%) of the bisector profile. One quantity is the bisector velocity span (BVS), which is the velocity difference between two flux levels with 25\% and 75\% of the bisector. Another is the bisector velocity curvature (BVC), which is the difference between two velocity spans in the upper half (between two flux levels with 50\% and 75\% of the bisector) and the lower half (25\% and 50\%). The other is the bisector velocity displacement (BVD), which is the average of the velocities at the three flux levels (25\%, 50\%, and 75\%). These bisector quantities for HD 100655 are shown in figure \ref{fig6}. The average values of BVS and BVC are 7.8 $\pm$ 8.1 m s$^{-1}$ and 2.4 $\pm$ 3.9 m s$^{-1}$, respectively. The BVS values may be increased due to rotational stellar spots, which may invoke photometric variation. However, since the average value of the BVS is one ninth of the velocity differences ($\sim$70 m s$^{-1}$) between the two templates, we consider both BVS and BVC value to be essentially zero, meaning that the cross-correlation profiles are symmetric. Moreover, the average value of the BVD ($-$70.1 $\pm$ 17.9 m s$^{-1}$) is consistent with the velocity difference between the two templates. Thus, the cause of the velocity difference is considered to be a parallel shift of spectral lines, not variations in spectral line shapes. Hence, the observed radial velocity variation of HD 100655 is best explained by the orbital motion of a planetary companion, not by intrinsic activity, such as rotational modulation and pulsation.
\section{Discussion}
We detected a planetary companion orbiting the clump giant star HD 100655 based on the precise Doppler spectroscopy survey conducted by the Korean$-$Japanese planet search program. The radial velocity variation of the star discovered during early observation at BOAO indicated the existence of a possible planetary companion, and the orbital parameters of the companion were determined by follow-up observations at BOAO and OAO. Adopting a mass of 2.4 $M_{\odot}$ for HD 100655, we found that the planetary companion has a minimum mass of 1.7 $M_{\mathrm{J}}$ and a semi-major axis of 0.76 AU. This is the lowest-mass planet among those discovered around giant stars with masses larger than 1.9 $M_{\odot}$. Fourteen planetary companions and six brown dwarf-mass companions have been detected so far around such giants by ongoing precise Doppler surveys, and the discoveries bring out some characteristic properties of the planetary systems.
Figure \ref{fig7} plots mass of substellar companions with semi-major axes less than 3 AU against host-star's mass (updated version of figure 5 of \cite{Omiya2009}). This figure includes intermediate-mass (1.5 $M_{\odot}$ $\leq$ $M$ $\leq$ 4 $M_{\odot}$) giants and subgiants ($filled$ $circles$), intermediate-mass dwarfs ($open$ $circles$), solar-mass stars ($M$ $<$ 1.5 $M_{\odot}$, $open$ $triangles$), and HD 100655 ($star$). Solid and dot-dashed lines indicate the lower-mass limits of companions detectable by current Doppler surveys for semi-major axes of 0.6 AU and 3 AU, respectively. The limits correspond to companion masses that give rise to semi-amplitudes of radial velocity variations of their host stars as large as three times the typical radial velocity jitters, which are 5 m s$^{-1}$ for subgiants (1.5$-$1.9 $M_{\odot}$) and 20 m s$^{-1}$ for clump giants (1.9$-$4 $M_{\odot}$) (\cite{Johnson2010c}, \cite{Sato2005}). Two unpopulated regions of substellar companions orbiting intermediate-mass subgiants and giants appear in regions (a) and (b) in figure \ref{fig7} \citep{Omiya2009}. The planet orbiting HD 100655 is located below the detection limit and the region (b) in figure \ref{fig7} because of its small root mean-square scatter of the residual radial velocities ($\sim$11 m s$^{-1}$), that is, its small radial velocity jitter. The existence of this planet suggests a possibility that low-mass giant planets can form around $\sim$2.4 $M_{\odot}$ stars, while this planet could possibly have a small orbital inclination, and thus a high actual mass. Therefore, a paucity of low-mass companions orbiting massive intermediate-mass giants, roughly indicated by the region (b), might partly be caused by an observational bias due to the high detection limit. In this respect, observational surveys more sensitive to lower-mass substellar companions are necessary.
The mass distribution of substellar companions orbiting 1.5$-$3 $M_{\odot}$ stars may also depend on the semi-major axes of the companions. Figure \ref{fig8} is a plot of semi-major axis of substellar companion versus host star mass. Crosses, circles and filled circles indicate brown dwarf-mass companions (13$-$30 $M_{\mathrm{J}}$), "superplanets" (6$-$13 $M_{\mathrm{J}}$), and normal giant planets (1$-$6 $M_{\mathrm{J}}$), respectively. Solid, dot-dashed and dotted lines indicate the typical farthest orbital distances of companions detectable by current Doppler surveys for companion masses of 3, 4 and 5 $M_{\mathrm{J}}$, respectively. The distances correspond to orbital semi-major axes that the companions induce radial velocity variations of their host stars with semi-amplitudes as large as three times the typical radial velocity jitter, which is 20 m s$^{-1}$ for clump giants (1.9$-$3 $M_{\odot}$) \citep{Sato2005}. In figure \ref{fig8}, some interesting properties of substellar companions are suggested in three stellar mass ranges. Almost all the planets orbiting 1.5$-$1.9 $M_{\odot}$ stars are normal giant planets, and are located on orbits with semi-major axes of $>$1 AU. Many planets orbiting 1.9$-$2.5 $M_{\odot}$ stars seems to be classified in two groups\footnote[3]{We note that $<$3 $M_{\mathrm{J}}$ ($<$5 $M_{\mathrm{J}}$) planets orbiting such stars at 1 AU (3 AU) are below the lower-mass limits for typical detectable planets.}: normal giant planets at inner orbits (0.6$-$1.3 AU) and superplanets at outer orbits (1.9$-$3 AU). HD 100655 b is included in the group of the normal giant planets. All planet-mass companions orbiting 2.5$-$3 $M_{\odot}$ stars reside at semi-major axes larger than 1.9 AU, while all brown dwarf-mass companions are orbiting at semi-major axes less than 1.9 AU. Although the number of known substellar companions discovered around stars with $>$2.5 $M_{\odot}$ is still small, the distribution of substellar companions around 1.9$-$2.5 $M_{\odot}$ stars may differ from those around 1.5$-$1.9 $M_{\odot}$ and 2.5$-$3 $M_{\odot}$ stars.
To reproduce the distribution of giant planets around 1.9$-$2.5 $M_{\odot}$ giant stars, two scenarios can be suggested. One is the planet engulfment scenario caused by stellar evolution of primary stars. Most of the host stars are clump giants that should have experienced the RGB phase, which triggers rapid stellar expansion. \citet{Villaver2009} suggested that the primary stars can preferentially capture more massive planetary companions by tidal interaction in the RGB phase. Thus, the superplanets with semi-major axes of $<$1.9 AU might have been preferentially engulfed by their primary stars even if they had existed, leaving normal giant planets. However, according to \citet{Kunitomo2011}, the critical semi-major axis, within which a primary star can engulf planetary companions, decreases from $\sim$1.5(0.4) AU for 1.7(2.0) $M_{\odot}$ stars to $\sim$0.2 AU for 2.1 $M_{\odot}$ stars and thus all the planets with semi-major axis larger than 0.6 AU around 2.0$-$2.5 $M_{\odot}$ stars can survive the RGB phase regardless of their masses. Therefore, the observational properties would not be quantitatively explained by only this mechanism.
The other scenario is that the distribution of the planets is primordially originated from planet migration in protoplanetary disk. Dependence of Type-II migration rate on planet mass may separate locations of low-mass giant planets and superplanets. For example, based on the equation (1) of \citet{Currie2009}, 2 $M_{\mathrm{J}}$ and 8 $M_{\mathrm{J}}$ planets that formed in circular orbits with a semi-major axis of 3.8 AU around 2 $M_{\odot}$ stars can migrate to inner orbits with semi-major axes of $\sim$0.7 AU and $\sim$2.8 AU, respectively, assuming a disk dissipation time of 1 Myr. Thus, the observed orbital distribution of planets around 1.9$-$2.5 $M_{\odot}$ stars may be explained by this mechanism, and if this is the case many undetected lower-mass planets should be expected at distances of 1$-$3 AU because giant planets can form at any distance beyond the snow line. It should be noticed, however, the observed semi-major axis distributions of planetary systems around 1.5$-$1.9 $M_{\odot}$ and 2.5$-$3 $M_{\odot}$ stars might not be explained by only the effect of the migration.
Additionally, Type-II migration may not be only the mechanism that can locate giant planets around intermediate-mass stars. The magnetorotational instability-dead zone in the protoplanetary disks may encourage formations of giant planets only at $\sim$1 AU around intermediate-mass stars \citep{Kretke2009}. In this case, a drop-off of giant planets at $>\sim$1 AU might exist around such stars. However, the distribution of $<$3 $M_{\mathrm{J}}$ planets at larger than 1 AU around 1.9$-$3 $M_{\odot}$ stars has not been clarified yet due to the detection limits of current planet searches.
Thus, in order to examine roles of these mechanisms on planet formation and evolution around intermediate-mass stars, it is required to evaluate the semi-major axes distribution by further Doppler surveys of intermediate-mass stars with masses of $>$1.9 $M_{\odot}$ sensitive to lower-mass planets.
\bigskip
This research was supported as a Korea-Japan Joint Research Project under the Japan-Korea Basic Scientific Cooperation Program between Korea Science and Engineering Foundation (KOSEF) and Japan Society for the Promotion of Science (JSPS). This research is based on data collected at Bohyunsan Optical Astronomy Observatory (BOAO) that is operated by Korea Astronomy and Space Science Institute (KASI) and Okayama Astrophysical Observatory (OAO) that is operated by National Astronomical Observatory of Japan (NAOJ). We gratefully acknowledge the support from the staff members of BOAO and OAO during the observations. BCL acknowledges the support from Engineering Foundation (KOSEF) through the Science Research Center (SRC) program. TSY acknowledges the support from Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (No. 2010-0023430). Data analyses were in part carried out on common use data analysis computer system at the Astronomy Data Center, ADC, of the National Astronomical Observatory of Japan. This research has made use of the SIMBAD database, operated at CDS, Stransbourg, France.
|
\section{Supplementary Material}
\label{supp}
\subsection{Steps of Algorithm and Comments}
\label{details}
Our quantum algorithm works by the following sequence of steps.
\begin{enumerate}
\item \textbf{Prepare the free vacuum.}
Improving upon the efficiency of earlier, more general, state-construction
methods \cite{Zalka, Grover_Rudolph}, Kitaev and Webb
developed a quantum algorithm for constructing multivariate Gaussian
superpositions \cite{Kitaev_Webb}.
For large $\mathcal{V}$, the dominant cost in Kitaev and Webb's
method for producing $\mathcal{V}$-dimensional multivariate
Gaussians is the computation of the $\mathbf{L}\mathbf{D}\mathbf{L}^T$
decomposition of the inverse covariance matrix, where $\mathbf{L}$ is a
unit lower triangular matrix, and $\mathbf{D}$ is a diagonal matrix.
This can be done in $\tilde{O}(\mathcal{V}^{2.376})$ time with established
classical methods~\cite{Bunch, Coppersmith}.
(The notation $f(n) = \tilde{O}(g(n))$ means $f(n) = O(g(n) \log^c(n))$
for some constant $c$.)
The computation of the matrix
elements of the covariance matrix itself is easy because, for large $V$,
the sum
\begin{equation}
\label{freepropagator}
G^{(0)}(\mathbf{x} - \mathbf{y}) = \sum_{\mathbf{p} \in \Gamma} L^{-d}
\frac{1}{2 \omega(\mathbf{p})} e^{i \mathbf{p} \cdot (\mathbf{x}_i -
\mathbf{x}_j)}
\end{equation}
defining the propagator of the lattice theory is well approximated
by an easily evaluated integral.
\item \textbf{Excite wavepackets.}
The span of $\ket{\mathrm{vac}(0)} \ket{0}$ and
$\ket{\psi} \ket{1}$ is an invariant subspace, on which $H_\psi$
acts as
\begin{eqnarray}
H_\psi \ket{\mathrm{vac}(0)} \ket{0} & = & \ket{\psi} \ket{1} \,, \\
H_\psi \ket{\psi} \ket{1} & = & \ket{\mathrm{vac}(0)} \ket{0} \,.
\end{eqnarray}
Thus,
\begin{equation}
e^{-i H_\psi \pi/2} \ket{\mathrm{vac}(0)} \ket{0} = -i \ket{\psi} \ket{1} \,.
\end{equation}
Hence, by simulating a time evolution according to the Hamiltonian
$H_\psi$, we obtain the desired wavepacket state $\ket{\psi}$, up to an
irrelevant global phase and extra qubit, which can be discarded. After
rewriting $H_\psi$ in terms of the operators $\phi(\mathbf{x})$ and
$\pi(\mathbf{x})$, one sees that simulating $H_\psi$ is a very similar task
to simulating $H$, and can be done with the same techniques.
The only errors introduced at this step are due to the finite separation
distance $\delta$ between wavepackets, and are of order $\epsilon \sim
e^{-\delta/m}$.
(However, our wavepackets have a constant spread in momentum, and thus
differ from the idealization of particles with precisely defined momenta.)
The wavepacket preparation thus has complexity scaling
linearly with $n_{\mathrm{in}}$, the number of particles being
prepared, and necessitates a dependence $V \sim n_{\mathrm{in}}
\log(1/\epsilon)$.
\item \textbf{Adiabatically turn on the interaction.}
For $0 \leq s \leq 1$, let
\begin{equation}
\label{HS}
H(s) = \sum_{\mathbf{x} \in \Omega} a^d \left[ \frac{1}{2} \pi(\mathbf{x})^2
+ \frac{1}{2} (\nabla_a \phi)^2 (\mathbf{x}) + \frac{1}{2} m_0^2(s)
\phi(\mathbf{x})^2 + \frac{\lambda_0(s)}{4!} \phi(\mathbf{x})^4 \right]
\end{equation}
with $\lambda_0(0) = 0$.
$U_j$ is the unitary time evolution induced by $H(t/\tau)$ from
$t=\frac{j\tau}{J}$ to $t=\frac{(j+1)\tau}{J}$, namely,
\begin{eqnarray}
U_j & = & T \left\{ \exp \left[ -i \int_{j/J}^{(j+1)/J} H(s) \tau ds
\right] \right\} \,,
\end{eqnarray}
where $T\{ \cdot \}$ indicates the time-ordered product.
We suppress the dynamical phases by choosing $J$ to be sufficiently large.
The choice of a suitable ``path'' $\lambda_0(s),m_0^2(s)$, and
the complexity of this state-preparation process depends in a
complicated manner on the parameters in $H$ (\sect{preparing}).
\item \textbf{Simulate Hamiltonian time evolution.}
\item \textbf{Adiabatically turn off the interaction.}
The adiabatic turn-off of the coupling is simply the
time-reversed version of the adiabatic turn-on.
\item \textbf{Measure occupation numbers of momentum modes.}
For a given $\mathbf{p}$, measurement of
$L^{-d} a_{\mathbf{p}}^\dag a_{\mathbf{p}}$ by phase estimation
can be implemented with $O \left( \mathcal{V}^{2+\frac{1}{2k}} \right)$
quantum gates via a $k\th$-order Suzuki-Trotter formula.
Furthermore, if we instead simulate localized detectors, the
computational cost becomes independent of $V$ (much as the
computational cost of creating local wavepackets is independent of
$V$), but the momentum resolution becomes lower, as dictated by the
uncertainty principle.
\end{enumerate}
The allowable rate of adiabatic increase of the coupling constant
during state preparation is determined by the physical mass of the
theory. In the weakly coupled case, this can be calculated
perturbatively. In the strongly coupled case, such a calculation is no
longer possible. Thus one is left with the
problem of determining how fast one can perform the adiabatic state
preparation without introducing errors. Fortunately, one can easily
calculate the mass on a quantum computer, as follows. First, one
adiabatically prepares the interacting vacuum state at some small
$\lambda_0$, and measures the energy of the vacuum using phase
estimation.
The speed at which to increase $\lambda_0$ can be chosen
perturbatively for this small value of $\lambda_0$. Next, one
adiabatically prepares the state with a single zero-momentum particle
at the same value of $\lambda_0$, and measures its energy using phase
estimation. Subtracting these values yields the physical mass. This
value of the physical mass provides guidance as to the speed of
adiabatic increase of the coupling to reach a slightly higher
$\lambda_0$. Repeating this process for successively higher
$\lambda_0$ allows one to reach strong coupling, while always having an
estimate of mass by which to choose a safe speed for adiabatic state
preparation. In addition, mapping out the physical mass as a function
of bare parameters (hence, for example, mapping out the phase diagram)
may be of independent interest.
\subsection{Efficiency}
\label{efficiency}
To quantify the precision of a simulation, we demand that the
probability of a given scattering event in the simulation differ from
the true physical probability by no more than $\pm \epsilon$. There
are various sources of error: discretization of space, Trotter approximations,
imperfect adiabaticity, discretization and cutoff of the field at each
site, and imperfect spatial separation of particles in the asymptotic in
and out states. In a theory with a non-zero mass, errors due to imperfect
particle separation shrink exponentially with distance.
Thus, $V$ needs to scale only logarithmically
with $\epsilon$. Similarly, by the analysis of \sect{qubits},
the number of qubits per site scales only logarithmically with
$\epsilon$. By \eq{hightrotter}, the errors resulting from use of a
$k\th$-order Suzuki-Trotter formula with $n$ timesteps are $\epsilon \sim
n^{-2k}$. Thus, the complexity scales as $\epsilon^{-1/2k}$. For
large $k$, the dominant contributions to scaling with $\epsilon$ are
spatial discretization and imperfect adiabaticity.
The effect of spatial discretization is captured by (infinitely many)
additional terms in the effective Hamiltonian. Truncation of
these terms alters the calculated probability
of scattering events. In particular, the two dominant extra terms in the
effective Hamiltonian are $\sum_i \phi \partial_i^4 \phi$ and $\phi^6$
terms, arising from discretization of $(\nabla_a \phi)^2$ and
quantum effects, respectively. The coefficient of the
$\sum_i \phi \partial_i^4 \phi$ term is $O(a^2)$, and the coefficient
of the $\phi^6$ term is $O(a^{5-d})$, so that the former dominates for
$d=1,2$, whereas the latter makes a comparable contribution for
$d=3$. Thus, the overall discretization error is
\begin{equation}
\label{spatial_contrib}
\epsilon = O(a^2) \,, \quad d=1,2,3 \,.
\end{equation}
(To improve the scaling, one can use better finite differences
to approximate the derivative, and/or include the $\phi^6$ operator.
However, renormalization and mixing of the coefficients make this idea
more complicated than it is in standard numerical analysis.)
The diabatic errors at weak coupling are estimated and summarized in
\sect{weak}. The errors are quantified by a probability $\epsilon$ of
observing stray particles. Substituting the $a \sim \sqrt{\epsilon}$
dependence from \eq{spatial_contrib} into \eq{Gstrict}
yields\footnote{Whether we use \eq{Gstrict} or \eq{Glenient}
affects only the scaling with $V$.}
\begin{equation}
G_{\mathrm{adiabatic}} \sim \left( \frac{1}{\epsilon}
\right)^{1+d/2+o(1)} \,,\quad d=1,2,3
\end{equation}
scaling for the adiabatic state preparation. We use little-$o$
notation to convey precisely that we are neglecting both logarithmic
factors and contributions to the exponent that become arbitrarily
small as we use higher-order Suzuki-Trotter formulae. The other slow
part of the algorithm is the preparation of the free vacuum. This
scales as
\begin{equation}
G_{\mathrm{prep}} = \tilde{O}(\mathcal{V}^{2.376}) = \tilde{O}(a^{-2.376d}) =
\tilde{O}(\epsilon^{-1.188d}) \,,
\end{equation}
where the last equality follows from \eq{spatial_contrib}.
Thus, in $d=1$ the adiabatic state preparation is the dominant cost,
whereas in $d=2,3$ the preparation of the free vacuum dominates. This leaves
a final asymptotic scaling of
\begin{equation}
G_{\mathrm{total}} = O(G_{\mathrm{adiabatic}} + G_{\mathrm{prep}}) =
\left\{ \begin{array}{ll}
\left( \frac{1}{\epsilon} \right)^{1.5+o(1)} \,, & d=1, \\
\left( \frac{1}{\epsilon} \right)^{2.376+o(1)} \,, & d=2, \\
\left( \frac{1}{\epsilon} \right)^{3.564+o(1)} \,, & d=3.
\end{array} \right.
\end{equation}
The number of quantum gates used to simulate the
strongly coupled theory has scaling in $1/(\lambda_c - \lambda_0)$ and
$p$ that is dominated by adiabatic state preparation (\sect{strong}).
We also estimate scaling with $n_{\mathrm{out}}$ as follows. For
two incoming particles with momenta $\mathbf{p}$ and $\mathbf{-p}$, the
maximum number of kinematically allowed outgoing particles is $n_{\mathrm{out}}
\sim p$. For continuum behavior, $p = \eta/a$ for constant $\eta \ll
1$. Furthermore, one needs $V \sim n_{\mathrm{out}}$ to obtain
good asymptotic out states separated by a distance of at least $\sim
1/m_0$. Thus, $\mathcal{V} \sim n_{\mathrm{out}}^{d+1}$, so one needs
$n_{\mathrm{out}}^{2.376 (d+1)}$ gates to prepare the free
vacuum and, by \eq{strongpscale}, $n_{\mathrm{out}}^{2d+3+o(1)}$ gates to
reach the interacting theory adiabatically. (The adiabatic
turn-off takes no longer than the adiabatic turn-on.)
Hence the total scaling in $n_{\mathrm{out}}$ is dominated by
preparation of the free vacuum in three-dimensional spacetime, but by
adiabatic turn-on in two-dimensional spacetime. These results are
summarized in Table~\ref{strongtable}.
\subsection{Mass Renormalization}
\label{QPT}
The physical, or renormalized, mass as a function of the coupling features
prominently in our calculations. For the weak-coupling regime, its form is
obtained by perturbation theory. For the strong-coupling
regime, we use its known behavior near the phase transition.
At first order in the coupling, the shift of the squared mass is given by
$i$ times the one-loop Feynman diagram
\begin{equation} \label{diag1}
\includegraphics[width=1.2in]{lineloop.eps} \,.
\end{equation}
At second order, there is also a contribution from the two-loop diagram
\begin{equation} \label{diag2}
\includegraphics[width=1.2in]{sunset.eps} \,.
\end{equation}
The calculation of these diagrams is quite analogous to standard
calculations in perturbative quantum field theory, but there
are a couple of differences. First, the propagator is different because
of the discretization. Secondly, integrals over components $1,\ldots, d$
(but not component $0$) of loop momenta are cut off by $\pi/a$, that is,
the lattice spacing acts as an ultraviolet regulator. These differences
alter the nature the integrals and hence what methods can be used to
evaluate them.
The existence of a phase transition in the $\phi^4$ theory in
$D=2$ or $3$ spacetime dimensions was shown rigorously
in \cite{Glimm:1974tz,Guerra:1975ym,McBryan:1976ga}.
As the system approaches it, thermodynamic functions and correlation
functions exhibit power-law behavior, as is characteristic of
a second-order phase transition. In particular, for constant $m_0^2$,
\begin{equation}
\label{numass}
m \sim |\lambda_0 - \lambda_c|^\nu \,,
\end{equation}
where $\lambda_c$, the critical value of the coupling, depends on $m_0^2$.
Empirically, it has been found that systems with second-order phase
transitions can be classified into universality classes.
Within each class, critical exponents are universal, taking the same
values for all systems.
(This universality is explained by the concept of the renormalization
group.) The $\phi^4$ theory is believed to be in the same
universality class as the Ising model, for which
\begin{equation}
\label{nu}
\nu = \left\{ \begin{array}{ll} 1 \,, & D=2 \,,
\\
0.63\ldots \,, & D=3 \,.
\end{array} \right.
\end{equation}
The value above for $D=3$ has also been obtained directly in the $\phi^4$
theory by Borel resummation \cite{LeGuillou:1977ju}.
In $D=4$ dimensions, in contrast, the believed triviality of the
continuum $\phi^4$ theory implies that there is no non-trivial fixed point
of the renormalization group and hence no phase transition as one varies
($m_0^2$, $\lambda_0$). Moreover, triviality places bounds on the
maximum value of the renormalized coupling \cite{Luscher:1987ay}.
In particular, strong coupling requires $p a$ to be $O(1)$:
in the continuum-like regime, renormalized perturbation theory
should be valid.
\subsection{Representation by Qubits}
\label{qubits}
The required number of qubits per site is
\begin{equation}
\label{placevalue}
n_b = \log \left( 1+2 \lfloor \phi_{\max}/\delta_{\phi} \rfloor \right)
\,.
\end{equation}
In this section we show that one can simulate processes at
energy scale $E$, while maintaining $1-\epsilon$ fidelity to the exact
state, with $n_b$ logarithmic in $1/a$, $1/\epsilon$, and $V$. Our
analysis is nonperturbative, and thus applies equally to strongly and
weakly coupled $\phi^4$ theory.
Let $\ket{\psi}$ be the state, expressed in the field representation,
namely,
\begin{equation}
\ket{\psi} = \int_{-\infty}^\infty d \phi_1 \ldots
\int_{-\infty}^\infty d \phi_{\mathcal{V}} \
\psi(\phi_1,\ldots,\phi_{\mathcal{V}})
\ket{\phi_1, \ldots, \phi_{\mathcal{V}}} \,,
\end{equation}
and let
\begin{equation}
\ket{\psi_{\mathrm{cut}}} = \int_{-\phi_{\max}}^{\phi_{\max}} d \phi_1
\ldots \int_{-\phi_{\max}}^{\phi_{\max}} d \phi_{\mathcal{V}} \
\psi(\phi_1, \ldots, \phi_{\mathcal{V}}) \ket{\phi_1, \ldots
\phi_{\mathcal{V}}} \,.
\end{equation}
Then
\begin{equation}
\braket{\psi}{\psi_{\mathrm{cut}}} = \int_{-\phi_{\max}}^{\phi_{\max}} d \phi_1
\ldots \int_{-\phi_{\max}}^{\phi_{\max}} d \phi_{\mathcal{V}}
\ \rho(\phi_1, \ldots, \phi_{\mathcal{V}}) \,,
\end{equation}
where $\rho$ is the probability distribution
\begin{equation}
\rho(\phi_1,\ldots, \phi_{\mathcal{V}}) = | \psi(\phi_1, \ldots,
\phi_{\mathcal{V}})|^2 \,.
\end{equation}
In other words, $\braket{\psi}{\psi_{\mathrm{cut}}} = 1 -
p_{\mathrm{out}}$, where $p_{\mathrm{out}}$ is the probability that at
least one of $\phi_1,\ldots,\phi_{\mathcal{V}}$ is out of the range
$[-\phi_{\max},\phi_{\max}]$. By the union bound
($ \mathrm{Pr}(A \cup B) \leq \mathrm{Pr}(A) + \mathrm{Pr}(B)$),
\begin{equation}
\braket{\psi}{\psi_{\mathrm{cut}}} \geq 1 - \mathcal{V}
\max_{\mathbf{x} \in \Omega} p_{\mathrm{out}}(\mathbf{x}) \,,
\end{equation}
where $p_{\mathrm{out}}(\mathrm{x})$ is the probability that
$\phi(\mathbf{x})$ is out of the range $[-\phi_{\max},\phi_{\max}]$.
Let $\mu_{\phi(\mathbf{x})}$ and $\sigma_{\phi(\mathbf{x})}$ denote
the mean and standard deviation of $\phi(\mathbf{x})$ determined by
$\rho$. By Chebyshev's inequality, choosing $\phi_{\max} =
\mu_{\phi(\mathbf{x})} + c \sigma_{\phi(\mathbf{x})}$ ensures
\begin{equation}
p_{\mathrm{out}}(\mathbf{x}) \leq \frac{1}{c^2} \,.
\end{equation}
Thus, choosing
\begin{equation}
\label{phichoice}
\phi_{\max} = O \left( \max_{\mathbf{x} \in \Omega} \left(
\mu_{\phi(\mathbf{x})} + \sqrt{\frac{\mathcal{V}}{\epsilon}}
\sigma_{\phi(\mathbf{x})} \right) \right)
\end{equation}
ensures $\braket{\psi}{\psi_{\mathrm{cut}}} \geq 1-\epsilon$.
Next, we observe the following, which is straightforward to prove.
\begin{proposition}
\label{canonfourier}
Let $\hat{p}$ and $\hat{q}$ be Hermitian operators on
$L^2(\mathbb{R})$ obeying the canonical commutation relation
$[\hat{p},\hat{q}]=i \mathds{1}$. Then the eigenbasis of $\hat{p}$ is the
Fourier transform of the eigenbasis of $\hat{q}$.
\end{proposition}
By Proposition~\ref{canonfourier}, the eigenbasis of $a^d \pi(\mathbf{x})$
is the Fourier transform of the eigenbasis of $\phi(\mathbf{x})$. Thus,
discretizing $\phi(\mathbf{x})$ in increments of
$\delta_{\phi(\mathbf{x})}$ is roughly equivalent to the truncation $-
\pi_{\max} \leq \pi(\mathbf{x}) \leq \pi_{\max}$, where
\begin{equation}
\label{pimax}
\pi_{\max} = \frac{1}{a^d \delta_{\phi(\mathbf{x})}} \,.
\end{equation}
By the same argument used to choose $\phi_{\max}$, choosing
\begin{equation}
\label{pichoice}
\pi_{\max} = O \left( \max_{\mathbf{x} \in \Omega} \left(
\mu_{\pi(\mathbf{x})} + \sigma_{\pi(\mathbf{x})}
\sqrt{\frac{\mathcal{V}}{\epsilon}} \right) \right)
\end{equation}
ensures fidelity $1-\epsilon$ between $\ket{\psi}$ and its truncated
and discretized version.
To obtain useful bounds on $\phi_{\max}$ and $\pi_{\max}$, we must bound
$\mu_{\phi(\mathbf{x})}$, $\sigma_{\phi(\mathbf{x})}$,
$\mu_{\pi(\mathbf{x})}$, and $\sigma_{\pi(\mathbf{x})}$.
To this end, we make the following straightforward observation.
\begin{proposition}
\label{moments}
Let $M$ be a Hermitian operator and let $\ket{\psi}$ be a quantum
state. Then $|\bra{\psi} M \ket{\psi}| \leq \sqrt{ \bra{\psi} M^2
\ket{\psi} }$.
\end{proposition}
\begin{proof}
For brevity, let $\langle Q \rangle = \bra{\psi} Q \ket{\psi}$ for any
observable $Q$. The operator $\left( M - \langle M \rangle \mathds{1}
\right)^2$ is positive semidefinite. Thus,
\begin{eqnarray}
0 & \leq & \left\langle \left( M - \langle M \rangle \mathds{1} \right)^2
\right\rangle \\
& = & \left\langle M^2 - 2 \langle M \rangle M + \langle M \rangle^2 \mathds{1}
\right\rangle \\
& = & \langle M^2 \rangle - \langle M \rangle^2 \,.
\end{eqnarray}
\end{proof}
\noindent
Applied to the definitions
\begin{eqnarray}
\mu_{\phi(\mathbf{x})} & = & \bra{\psi} \phi(\mathbf{x}) \ket{\psi} \,, \\
\sigma_{\phi(\mathbf{x})} & = & \sqrt{ \bra{\psi} \phi(\mathbf{x})^2
\ket{\psi} - \bra{\psi} \phi(\mathbf{x}) \ket{\psi}^2} \,, \\
\mu_{\pi(\mathbf{x})} & = & \bra{\psi} \pi(\mathbf{x}) \ket{\psi} \,, \\
\sigma_{\pi(\mathbf{x})} & = & \sqrt{ \bra{\psi} \pi(\mathbf{x})^2
\ket{\psi} - \bra{\psi} \pi(\mathbf{x}) \ket{\psi}^2} \,,
\end{eqnarray}
Proposition~\ref{moments} implies that $\mu_{\phi(\mathbf{x})}$ and
$\sigma_{\phi(\mathbf{x})}$ are each at most
$\sqrt{\bra{\psi} \phi(\mathbf{x})^2 \ket{\psi}}$, and
$\mu_{\pi(\mathbf{x})}$ and $\sigma_{\pi(\mathbf{x})}$ are each at
most $\sqrt{\bra{\psi} \pi(\mathbf{x})^2 \ket{\psi}}$. Thus, by
\eq{phichoice} and \eq{pichoice},
\begin{eqnarray}
\phi_{\max} & = & O \left( \max_{\mathbf{x} \in \Omega} \sqrt{
\frac{\mathcal{V}}{\epsilon} \bra{\psi} \phi(\mathbf{x})^2
\ket{\psi}} \right) \,,\\
\pi_{\max} & = & O \left( \max_{\mathbf{x} \in \Omega} \sqrt{
\frac{\mathcal{V}}{\epsilon} \bra{\psi} \pi(\mathbf{x})^2
\ket{\psi}} \right) \,,
\end{eqnarray}
so that, by \eq{placevalue} and \eq{pimax},
\begin{equation}
n_b = O \left( \log \left( a^d \frac{\mathcal{V}}{\epsilon}
\max_{\mathbf{x},\mathbf{y} \in \Omega} \sqrt{\bra{\psi} \pi(\mathbf{x})^2
\ket{\psi} \bra{\psi} \phi(\mathbf{y})^2 \ket{\psi}} \right) \right) \,.
\end{equation}
To establish logarithmic scaling of $n_b$, we need only prove
polynomial upper bounds on $\bra{\psi} \phi(\mathbf{x})^2 \ket{\psi}$
and $\bra{\psi} \pi(\mathbf{x})^2 \ket{\psi}$. Rather than making a physical
estimate of these expectation values, we prove simple
upper bounds that are probably quite loose. In the adiabatic state
preparation described in \sect{preparing}, the parameters
$m_0^2$ and $\lambda_0$ are varied. The following two propositions
cover all the combinations of parameters used in the adiabatic
preparation and subsequent scattering of both strongly and weakly
coupled wavepackets.
\begin{proposition}
\label{mpos}
Let $H$ be of the form shown in \eq{HS}. Suppose $m_0^2 > 0$ and
$\lambda_0 \geq 0$. Let $\ket{\psi}$ be any state of the field such
that $\bra{\psi} H \ket{\psi} \leq E$. Then $\forall \mathbf{x} \in
\Omega$,
\begin{eqnarray}
\label{phiboundpos}
\bra{\psi} \phi(\mathbf{x})^2 \ket{\psi} & \leq & \frac{2E}{a^d m_0^2} \,,\\
\label{piboundpos}
\bra{\psi} \pi(\mathbf{x})^2 \ket{\psi} & \leq & \frac{2E}{a^d} \,.
\end{eqnarray}
\end{proposition}
\begin{proof}
\begin{eqnarray}
E & \geq & \bra{\psi} H \ket{\psi} \\
& = &
\label{almost}
\bra{\psi} \sum_{\mathbf{x} \in \Omega} a^d \left[ \frac{1}{2}
\pi(\mathbf{x})^2 + \frac{1}{2} (\nabla_a \phi)^2(\mathbf{x}) +
\frac{m_0^2}{2} \phi(\mathbf{x})^2 + \frac{\lambda_0}{4!}
\phi(\mathbf{x})^2 \right] \ket{\psi} \\
& \geq & \bra{\psi} a^d \frac{m_0^2}{2} \phi(\mathbf{x})^2 \ket{\psi},
\end{eqnarray}
where the last inequality follows because all of the operators we have
dropped are positive semidefinite. This establishes
\eq{phiboundpos}. Similarly, we can drop all but the
$\pi(\mathbf{x})$ term from the right-hand side of \eq{almost},
leaving
\begin{equation}
E \geq \bra{\psi} a^d \frac{1}{2} \pi(\mathbf{x})^2 \ket{\psi} \,,
\end{equation}
which establishes \eq{piboundpos}.
\end{proof}
\begin{proposition}
\label{mneg}
Let $H$ be of the form shown in \eq{HS}. Suppose $m_0^2 \leq 0$ and
$\lambda_0 > 0$. Let $\ket{\psi}$ be any state of the field such that
$\bra{\psi} H \ket{\psi} \leq E$. Then $\forall \mathbf{x} \in
\Omega$,
\begin{eqnarray}
\label{phiboundneg}
\bra{\psi} \phi(\mathbf{x})^2 \ket{\psi} & \leq &
-\frac{24m_0^2}{\lambda_0} + \sqrt{\frac{36
m_0^4}{\lambda_0^2}+\frac{24}{\lambda_0 a^d} \left(
E + \frac{3(V-a^d)m_0^4}{2\lambda_0} \right)} \,,
\\
\label{piboundneg}
\bra{\psi} \pi(\mathbf{x})^2 \ket{\psi} & \leq & \frac{2}{a^d} \left(
E + \frac{3V m_0^4}{2 \lambda_0} \right) \,,
\end{eqnarray}
where $V$ is the physical volume.
\end{proposition}
\begin{proof}
The operator
\begin{equation}
U(\mathbf{x}) = \frac{m_0^2}{2} \phi(\mathbf{x})^2 +
\frac{\lambda_0}{4!} \phi(\mathbf{x})^4
\end{equation}
is sufficiently simple that we can directly calculate its minimal
eigenvalue $U_{\min}$. If $m_0^2 \leq 0$ and $\lambda > 0$, then
\begin{equation}
\label{Vmin}
U_{\min} = - \frac{3 m_0^4}{2 \lambda_0} \,.
\end{equation}
Thus, for \emph{any} state $\ket{\psi}$,
\begin{equation}
\label{anystate}
\bra{\psi} \sum_{\mathbf{x} \in \Omega} a^d U(\mathbf{x}) \ket{\psi}
\geq \frac{-3Vm_0^4}{2 \lambda_0} \,.
\end{equation}
Hence, recalling \eq{HS}, we obtain
\begin{eqnarray}
E & \geq & \bra{\psi} H \ket{\psi} \\
& = &
\label{beginning}
\bra{\psi} \sum_{\mathbf{x} \in \Omega} a^d \left[ \frac{1}{2}
\pi(\mathbf{x})^2 + \frac{1}{2} ( \nabla_a \phi)^2(\mathbf{x}) +
\frac{m_0^2}{2} \phi(\mathbf{x})^2 + \frac{\lambda_0}{4!}
\phi(\mathbf{x})^4 \right] \ket{\psi} \\
& \geq &
\label{secondtolast}
\bra{\psi} \sum_{\mathbf{x} \in \Omega} a^d \left[
\frac{1}{2} \pi(\mathbf{x})^2 + \frac{1}{2} (\nabla_a
\phi)^2(\mathbf{x}) \right] \ket{\psi} - \frac{3V m_0^4}{2 \lambda_0}
\\
& \geq &
\label{last}
\bra{\psi} \frac{a^d}{2} \pi(\mathbf{x})^2 \ket{\psi} -
\frac{3Vm_0^4}{2 \lambda_0} \,.
\end{eqnarray}
\eq{secondtolast} follows from \eq{anystate}. \eq{last} holds
(for any choice of $\mathbf{x}$) because all of the operators we have
dropped are positive semidefinite. This establishes
\eq{piboundneg}.
Similarly, dropping positive operators from \eq{beginning} and
using \eq{anystate} yield, for any $\mathbf{x}$,
\begin{equation}
a^d \bra{\psi} \left( \frac{m_0^2}{2} \phi(\mathbf{x})^2 +
\frac{\lambda_0}{4!} \phi(\mathbf{x})^4 \right) \ket{\psi} \leq
\left( E + \frac{3(V-a^d)m_0^4}{2\lambda_0} \right) \,.
\end{equation}
Applying Proposition \ref{moments} with $M = \phi(\mathbf{x})^2$ shows
that $\bra{\psi} \phi(\mathbf{x})^4 \ket{\psi} \geq \bra{\psi}
\phi(\mathbf{x})^2 \ket{\psi}^2$. Thus,
\begin{equation}
a^d \left[ \frac{m_0^2}{2} \bra{\psi} \phi(\mathbf{x})^2 \ket{\psi} +
\frac{\lambda_0}{4!} \bra{\psi} \phi(\mathbf{x})^2 \ket{\psi}^2
\right] \leq \left( E + \frac{3(V-a^d)m_0^4}{2\lambda_0} \right) \,.
\end{equation}
Via the quadratic formula, this implies \eq{phiboundneg}.
\end{proof}
\subsection{Adiabatic Preparation of Interacting Wavepackets}
\label{preparing}
In this section, we analyze the adiabatic state-preparation procedure.
To analyze the error due to finite
$\tau$ and $J$, we consider the process of preparing a single-particle
wavepacket. The procedure
performs similarly in preparing wavepackets for multiple particles
provided the particles are separated by more than the characteristic
length $1/m$ of the interaction.
The phase induced by $M_j$ on the momentum-$p$ eigenstate of $H(s)$
(with energy $E_p(s)$) is
\begin{equation}
\theta_j(p) = \left( E_p \left( \frac{j+1}{J} \right) + E_p \left(
\frac{j}{J} \right) \right) \frac{\tau}{2J} - \tau
\int_{j/J}^{(j+1)/J} ds E_p(s) \,.
\end{equation}
Taylor expanding $E_p$ about $s=(j+\frac{1}{2})/J$ yields
\begin{equation}
\label{thetaj}
\theta_j(p) = \frac{\tau}{12 J^3} \frac{\partial^2 E_p}{\partial s^2}
+ O(J^{-5}) \,.
\end{equation}
Thus the total phase induced is
\begin{eqnarray}
\theta(p) & = & \sum_{j=0}^{J-1} \theta_j(p) \\
& \simeq & \frac{\tau}{12 J^2} \int_0^1 ds \frac{\partial^2
E_p}{\partial s^2} \\
& = & \frac{\tau}{12 J^2} \left. \frac{\partial E_p}{\partial s}
\right|_0^1 \,, \label{thetap1}
\end{eqnarray}
where the approximation holds for large $J$. For a Lorentz-invariant
theory, $E_p(s)$ must take the form
\begin{equation}
\label{LI}
E_p(s) = \sqrt{p^2+m^2(s)} \,.
\end{equation}
This should be a good approximation for the lattice theory provided
the particle momentum satisfies $p \ll 1/a$. Substituting \eq{LI}
into \eq{thetap1} yields
\begin{equation}
\label{thetap2}
\theta(p) \simeq \frac{\tau}{24 J^2} \left. \frac{\frac{\partial
m^2}{\partial{s}}}{\sqrt{p^2 + m^2(s)}} \right|_0^1 \,.
\end{equation}
Next, we consider the effect of this phase shift on a wavepacket
centered around momentum $\bar{p}$. If the wavepacket is narrowly
concentrated in momentum, then we can Taylor expand $\theta(p)$ to
first order about $\bar{p}$:
\begin{equation}
\theta(p) \simeq \theta(\bar{p}) + \mathcal{D} \cdot (p-\bar{p}) \,,
\end{equation}
where
\begin{equation}
\mathcal{D} = \left. \frac{\partial \theta}{\partial p}
\right|_{\bar{p}} \,. \label{D}
\end{equation}
The phase shift $e^{i \mathcal{D} \cdot (p - \bar{p})}$ induces a
translation (in position space) of any wavepacket by a distance $\mathcal{D}$.
(The second-order term in the Taylor expansion induces
broadening.) From \eq{D} and \eq{thetap2}, we have
\begin{equation}
\label{phasecrit}
\mathcal{D} \simeq \left| \frac{\tau |\bar{p}| }{24 J^2}
\left. \frac{\frac{\partial m^2} {\partial s}}{\left( \bar{p}^2 +
m^2(s) \right)^{3/2}} \right|_{s=0}^{s=1} \right| \,.
\end{equation}
We next determine the complexity by demanding that the propagation length
$\mathcal{D}$ be restricted to some small constant, and that the
probability of diabatic particle creation be small. Together, these
criteria determine $J$ and $\tau$. We can obtain a tighter bound in
the perturbative case than in the general case, so we treat these
separately.
\begin{figure}
\begin{center}
\includegraphics[width=0.33\textwidth]{paths2.eps}
\caption{\label{paths} The dashed line illustrates schematically the
location of a quantum phase transition of $\phi^4$ theory in two and
three spacetime dimensions. A and B denote weakly and strongly coupled
continuum-like theories, respectively. We prepare them adiabatically
by following the arrows starting from the massive free theory
($m_0^2 >0$, $\lambda_0 = 0$). To maintain adiabaticity, the path
must not cross the quantum phase transition.}
\end{center}
\end{figure}
\subsubsection{Weak Coupling}
\label{weak}
In the perturbative continuum limit $a \to 0$, $m_0^2$ is
negative. For fixed small $a$, we can adiabatically approach a
perturbative continuum-like theory by taking the straight-line path
depicted in Fig.~\ref{paths}, namely, the following parameterization of
\eq{HS}:
\begin{eqnarray}
m_0^2(s) & = & (m^{(1)})^2 + s \lambda_0 \mu \,, \nonumber \\
\lambda_0(s) & = & s \lambda_0 \label{weakpath} \,.
\end{eqnarray}
Using perturbation theory (see diagram~\ref{diag1}), one finds that it
is particularly efficient to choose
\begin{equation}
\mu =
\left\{
\begin{array}{ll}
-\frac{1}{8\pi} \log\Big(\frac{64}{m^2a^2}\Big)
+ \cdots \,,
& d=1 \,,\\
[5pt]
-\frac{r_0^{(2)}}{16\pi^2}\frac{1}{a}
+ \cdots \,,
& d=2 \,,\\
[5pt]
-\frac{r_0^{(3)}}{32\pi^3}\frac{1}{a^2}
+ \cdots \,,
& d=3 \,,
\end{array}
\right.
\label{eq:path}
\end{equation}
so that, at first order in $\lambda_0$, the physical mass remains
fixed at $m^{(1)}$ for all $s$.
Here, $r_0^{(2)} = 25.379\ldots $ and $r_0^{(3)} = 112.948\ldots $.
In the perturbative regime, this
should ensure that the path does not cross the quantum phase
transition.
To calculate the variation of physical mass with $s$, we must go to
second order in $\lambda_0$ (see diagram~\ref{diag2}). The result is
\begin{equation}
\label{uptosecond}
m^2(s) = (m^{(1)} )^2 + s^2 m_2^2 + O(\lambda_0^3)\,,
\end{equation}
where
\begin{equation}
\label{m22}
m_2^2 = \left\{ \begin{array}{ll} O \left( \lambda_0^2/
(m^{(1)})^2 \right) \,, &
d=1 \,,
\vspace{4pt}\\
O\left(\lambda_0^2 \log (m^{(1)}a)\right) \,, &
d=2 \,,
\vspace{4pt}\\
O(\lambda_0^2/a^2) \,, &
d=3 \,.
\end{array} \right.
\end{equation}
Substituting \eq{uptosecond} into \eq{phasecrit} yields
\begin{equation}
\label{phasecrit2}
\frac{\tau |\bar{p}|}{12 J^2} \frac{m_2^2}{\left( \bar{p}^2 +
(m^{(1)})^2 + m_2^2 \right)^{3/2} } \leq \mathcal{D} \,.
\end{equation}
If we are considering a fixed physical process and using successively
smaller $a$ to achieve higher precision then, by \eq{m22},
it suffices to choose $J$ to scale as
\begin{equation}
\label{Jscale}
J = \left\{ \begin{array}{ll}
\tilde{O} \left(\sqrt{\frac{m^{(1)} \tau}{\lambda_0 \mathcal{D}}}
\right) \,, & d=1 \,, \vspace{5pt} \\
\tilde{O} \left( \sqrt{\frac{\tau}{\lambda_0 \mathcal{D}}} \right) \,,
& d=2 \,, \vspace{5pt} \\
\tilde{O} \left( \sqrt{\frac{a \tau}{\lambda_0 \mathcal{D}}}
\right) \,, & d = 3 \,. \end{array} \right.
\end{equation}
Note that, for $d=3$, $J$ is suppressed by $\sqrt{a}$. This is because,
as $s$ increases, the (uncancelled) two-loop contribution to the physical
mass makes the particle very heavy until $s$ is very close to one.
Hence, the particle propagates slowly, and less backward evolution is
required.
To determine $\tau$, we next consider adiabaticity. Let $H(s)$ be any
Hamiltonian differentiable with respect to $s$. Let
$\ket{\phi_l(s)}$ be an eigenstate $H(s) \ket{\phi_l(s)} = E_l(s)$
separated by a non-zero energy gap for all $s$. Let $\ket{\psi_l(t)}$
be the state obtained by Schr\"odinger time evolution according to
$H(t/\tau)$ with initial condition $\ket{\psi_l(0)} =
\ket{\phi_l(0)}$. The diabatic transition amplitude to any other eigenstate
$H(s) \ket{\phi_k(s)} = E_k(s)
\ket{\phi_k(s)}$ ($k \neq l$) is \cite{Messiah}
\begin{equation}
\label{traditional_integral}
\braket{\phi_k(s)}{\psi_l(\tau s)} \sim \int_{0}^{s} d \sigma
\frac{\bra{\phi_k(\sigma)} \frac{dH}{ds} \ket{\phi_l(\sigma)}}{E_l(\sigma)-E_k(\sigma)}
e^{i \tau (\varphi_k(\sigma)-\varphi_l(\sigma))} \left( 1 + O(1/\tau)
\right) \,.
\end{equation}
(The integrand is made well-defined by the phase convention
$\bra{\phi_k} \frac{d \ket{\phi_k}}{ds} = 0$.) Here,
\begin{equation}
\varphi_l(s) = \int_{0}^s d \sigma E_l(\sigma) \,.
\end{equation}
In the case that $E_l$, $E_k$, and $\bra{\phi_k} \frac{dH}{ds}
\ket{\phi_l}$ are $s$-independent, this integral gives
\begin{equation}
\label{traditional}
\braket{\phi_k(s)}{\psi_l(\tau s)} \sim \left( 1 - e^{i \tau
(E_k-E_l)s} \right) \frac{\bra{\phi_k}
\frac{dH}{ds} \ket{\phi_l}}{-i \tau (E_k-E_l)^2} (1 + O(1/\tau^2)) \,.
\end{equation}
In the case that these quantities are approximately $s$-independent,
\eq{traditional} should hold as an approximation.
In reality, we wish to prepare a wavepacket state, not an
eigenstate. However, the wavepacket is well separated from other
particles and narrowly concentrated in momentum space. Thus, we shall
approximate it as an eigenstate $\ket{\phi_l(s)}$. Furthermore, by our
choice of path, the energy gap is kept constant to first order in the
coupling, and thus \eq{traditional} should be a good approximation
to \eq{traditional_integral}.
Summing the transition amplitudes to some state $\ket{\phi_k}$ from
the $J$ steps in in our preparation process, and applying the
triangle inequality\footnote{The $O(J)$ scaling obtained by the
triangle inequality can be confirmed by a more detailed calculation
taking into account the relative phases of the contributions to the
total transition amplitude.} yield the following:
\begin{equation}
\label{totaldiabatic}
\left| \braket{\phi_k}{\psi_l(\tau)} \right| = O \left(
\frac{1}{\tau} \sum_{j=0}^J \left| \frac{\bra{\phi_k(j/J)} \frac{dH}{ds}
\ket{\phi_l(j/J)}}{(E_k(j/J) - E_l(j/J))^2} \right| \right) \,.
\end{equation}
The $j=0$ term in this sum can be evaluated exactly, because it arises
from the free theory. At $j \neq 0$ the theory is no longer exactly
solvable. However, one obtains the lowest-order contribution to the matrix
element $\bra{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3,\mathbf{p}_4;s=1}
\phi^4 \ket{\mathrm{vac}(1)}$ in renormalized perturbation theory
simply by taking the $j=0$ expression and replacing $m_0$
with the physical mass and $\lambda_0$ with the physical coupling. Our
adiabatic path \eq{weakpath} is designed so that the physical mass
at $s=1$ matches the bare mass at $j=0$ (at least to first order in
$\lambda_0$). Furthermore, the physical coupling differs from the bare
coupling only by a logarithmically divergent (in $a$) correction for
$d=3$ and non-divergent corrections for $d=1,2$. Thus we can make the
following approximation:
\begin{equation}
\label{summedadiabatic}
\left| \braket{\phi_k}{\psi_l(\tau)} \right| = \tilde{O} \left(
\frac{J}{\tau} \left| \frac{\bra{\phi_k(0)} \frac{dH}{ds}
\ket{\phi_l(0)}}{(E_k(0) - E_l(0))^2} \right| \right) \,.
\end{equation}
Diabatic errors come in two types, creation of particles from the
vacuum, and splitting of the incoming particles. The matrix element in the
numerator of \eq{summedadiabatic} can correspondingly be decomposed
as the sum of two contributions. We first consider particle creation
from the vacuum, approximating $\ket{\phi_j(s)}$ as
$\ket{\mathrm{vac}(s)}$.
By \eq{weakpath},
\begin{equation}
\frac{dH}{ds} = \sum_{\mathbf{x} \in \Omega} a^d \left[
\frac{\lambda_0}{4!} \phi^4(\mathbf{x}) + \lambda_0 \mu
\phi^2(\mathbf{x}) \right] \,.
\end{equation}
Substituting this into the numerator of \eq{summedadiabatic},
setting $\ket{\phi_l(0)} = \ket{\mathrm{vac}(0)}$, and expanding
$\phi$ in terms of creation and annihilation operators show that the
only potentially non-zero transition amplitudes are to states
$\ket{\phi_k(0)}$ of two or four particles. The transition amplitude
to states of four particles arise solely from the $\phi^4$ term in
$\frac{dH}{ds}$. The transition amplitude to states of two particles
has contributions from the $\phi^4$ term and the $\phi^2$ term in
$\frac{dH}{ds}$. These actually cancel, because of our choice of $\mu$.
(Note that this requires tuning of $\mu$.)
At $s=0$, the numerator of \eq{summedadiabatic} is therefore the
following:
\begin{equation}
\label{treelevel}
\bra{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3,\mathbf{p}_4}
\frac{\lambda_0}{4!}
\sum_{\mathbf{x} \in \Omega} a^d \phi^4(\mathbf{x})
\ket{\mathrm{vac}(0)} = \frac{ \lambda_0
\delta_{\mathbf{p}_1+\mathbf{p}_2+\mathbf{p}_3+\mathbf{p}_4,0}}
{4 V \sqrt{\omega(\mathbf{p}_1) \omega(\mathbf{p}_2)
\omega(\mathbf{p}_3) \omega(\mathbf{p}_4)}} \,.
\end{equation}
We obtain the probability of excitation due to creation of four
particles from the vacuum by squaring the amplitude estimated above,
and then summing over all allowed combinations of the four outgoing
momenta:
\begin{equation}
P_{\mathrm{create}} \sim
\sum_{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3,\mathbf{p}_4 \in \Gamma}
\frac{J^2 \lambda_0^2
\delta_{\mathbf{p}_1+\mathbf{p}_2+\mathbf{p}_3+\mathbf{p}_4,0}}
{V^2 \tau^2 (\omega(\mathbf{p}_1)+\omega(\mathbf{p}_2) +
\omega(\mathbf{p}_3) + \omega(\mathbf{p}_4))^4 \omega(\mathbf{p}_1)
\omega(\mathbf{p}_2) \omega(\mathbf{p}_3) \omega(\mathbf{p}_4)} \,.
\end{equation}
This sum is difficult to evaluate exactly; instead, we shall simply
estimate its asymptotic scaling. The question is, with which
parameter should we consider scaling?
There are at least three regimes in which classical methods for computing
scattering amplitudes break down or are inefficient: strong coupling,
large numbers of external particles, and high precision.
In this section we are considering only weak coupling
(that is, $\lambda/m^{4-D} \ll 1$), leaving discussion of strong coupling
until the next section.
For an asymptotically large number of external particles, the efficiency
of our algorithm depends upon strong coupling, for the following reason.
A connected Feynman diagram involving $n$ external particles must have
at least $v=O(n)$ vertices, so the amplitude for such a process
is suppressed by a factor of $\left( \frac{\lambda}{E^{4-D}}
\right)^v$, where $E$ is the energy scale of the process. Since
$E \geq m$, many-particle scattering events are exponentially rare
at weak coupling, and thus cannot be efficiently observed in
experiments or simulations. This leaves the high-precision
frontier. Recall that the perturbation series used in quantum field
theory are asymptotic but not convergent. Thus, perturbative methods
cannot be extended to arbitrarily high precision.
Hence, in this section we consider the quantum gate complexity of
achieving arbitrarily high precision.
To do so, one chooses $a$ small to obtain
small discretization errors, $V$ large to obtain better
particle separation, $\tau$ long to improve adiabaticity, and $J$
large enough to limit unwanted particle propagation as the interaction
is turned on. Thus, we wish to know the scaling
of $P_{\mathrm{create}}$ with $a$, $\tau$, $V$, and $J$. In this
context, we consider
$m$, $\lambda$, and $|\mathbf{p}_1|$ to be constants.
We now estimate the scaling of $P_{\mathrm{create}}$ as $a \to 0$.
\begin{eqnarray}
P_{\mathrm{create}} & \sim & \frac{J^2}{V^2 \tau^2}
\sum_{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3 \in \Gamma} \nonumber \\
& & \frac{\lambda_0^2}
{(\omega(\mathbf{p}_1)+\omega(\mathbf{p}_2)+\omega(\mathbf{p}_3) +
\omega(-\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3))^4
\omega(\mathbf{p}_1) \omega(\mathbf{p}_2) \omega(\mathbf{p}_3)
\omega(-\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3)} \nonumber \\
& \simeq & \frac{3 J^2}{V^2 \tau^2}
\sum_{\substack{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3 \in \Gamma \\ |\mathbf{p}_1| > |\mathbf{p}_2|,|\mathbf{p}_3|}} \nonumber \\
& &
\frac{\lambda_0^2}
{(\omega(\mathbf{p}_1)+\omega(\mathbf{p}_2)+\omega(\mathbf{p}_3)+\omega(-\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3))^4
\omega(\mathbf{p}_1) \omega(\mathbf{p}_2) \omega(\mathbf{p}_3) \omega(-\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3)} \nonumber \\
& \sim & \frac{J^2}{V^2 \tau^2}
\sum_{\substack{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3 \in \Gamma \\ |\mathbf{p}_1| > |\mathbf{p}_2|,|\mathbf{p}_3|}}
\frac{\lambda_0^2}
{\omega(\mathbf{p}_1)^6 \omega(\mathbf{p}_2) \omega(\mathbf{p}_3)} \nonumber \\
& \leq & \frac{J^2}{V^2 \tau^2}
\sum_{\mathbf{p}_1,\mathbf{p}_2,\mathbf{p}_3 \in \Gamma}
\frac{\lambda_0^2}
{\omega(\mathbf{p}_1)^6 \omega(\mathbf{p}_2) \omega(\mathbf{p}_3)} \nonumber \\
& \simeq & \frac{V J^2}{\tau^2} \int_{\Gamma} d^d p_1 \int_{\Gamma}
d^d p_2 \int_{\Gamma} d^d p_3 \frac{\lambda_0^2}{\omega(\mathbf{p}_1)^6
\omega(\mathbf{p}_2) \omega(\mathbf{p}_3)} \nonumber \\
& = &
\label{pcreate}
\left\{ \begin{array}{ll} \tilde{O} \left(
\frac{V J^2}{\tau^2} \right) \,, & d=1,2 \,, \vspace{5pt} \\
\tilde{O} \left( \frac{V J^2}{\tau^2 a} \right) \,, & d=3 \,.
\end{array} \right.
\end{eqnarray}
By \eq{Jscale} and \eq{pcreate},
\begin{equation}
P_{\mathrm{create}}= \begin{array}{ll}
\tilde{O} \left( \frac{V}{\tau} \right) \,, & d=1,2,3 \,.
\end{array}
\end{equation}
Next, we consider the process in which the time dependence of the
$\phi^4$ term causes a single particle to split into three. For this
process, the relevant matrix element is
\begin{equation}
\bra{\mathbf{p}_2,\mathbf{p}_3,\mathbf{p}_4} \frac{\lambda_0}{4 !}
\sum_{\mathbf{x} \in \Omega}
a^d \phi^4(\mathbf{x}) \ket{\mathbf{p}_1} = \frac{ \lambda_0
\delta_{\mathbf{p}_2+\mathbf{p}_3+\mathbf{p}_4,\mathbf{p}_1}}
{4 V \sqrt{\omega(\mathbf{p}_1) \omega(\mathbf{p}_2)
\omega(\mathbf{p}_3) \omega(\mathbf{p}_4)}} \,,
\end{equation}
where $\mathbf{p}_1$ is the momentum of the incoming particle.
By our choice of path, the physical mass is $s$-independent to
first order in the coupling, and the $s$ dependence of the
coupling is only logarithmically divergent as $a \to 0$. Thus,
by \eq{totaldiabatic},
\begin{equation}
\label{Psplitdef}
P_{\mathrm{split}} \sim \frac{J^2}{\tau^2 V^2}
\sum_{\mathbf{p}_2,\mathbf{p}_3,\mathbf{p}_4 \in \Gamma}
\frac{\lambda_0^2 \delta_{\mathbf{p}_2+\mathbf{p}_3+\mathbf{p}_4,\mathbf{p}_1}}
{(\omega(\mathbf{p}_2)+\omega(\mathbf{p}_3)+\omega(\mathbf{p}_4)-\omega(\mathbf{p}_1))^4
\omega(\mathbf{p}_1) \omega(\mathbf{p}_2) \omega(\mathbf{p}_3) \omega(\mathbf{p}_4)} \,.
\end{equation}
Let us now examine the divergence structure of $P_{\mathrm{split}}$ as
$a \to 0$. In the limit of large volume, the sum converges to the
following integral:
\begin{equation}
\frac{2J^2}{\tau^2} \int_\Gamma d^d p_2
\int_\Gamma d^d p_3 \frac{\lambda_0^2}{(\omega(\mathbf{p}_2)+\omega(\mathbf{p}_3) +
\omega(\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3) - \omega(\mathbf{p}_1))^4 \omega(\mathbf{p}_1) \omega(\mathbf{p}_2)
\omega(\mathbf{p}_3) \omega(\mathbf{p}_1-\mathbf{p}_2-\mathbf{p}_3)} \,. \\
\end{equation}
If this were divergent as $a \to 0$, then by approximating the integrand
with its value at large
$|\mathbf{p}_2|$ and $|\mathbf{p}_3|$, we would be able to isolate the
divergence:
\begin{equation}
P_{\mathrm{split}} \sim \frac{J^2 \lambda_0^2}{\tau^2
\omega(\mathbf{p}_1)} \int_\Gamma d^d p_2 \int_\Gamma d^d p_3
\frac{1}{(|\mathbf{p}_2|+|\mathbf{p}_3|+|\mathbf{p}_2+\mathbf{p}_3|)^4
|\mathbf{p}_2| |\mathbf{p}_3| |\mathbf{p}_2+\mathbf{p}_3|} \,.
\end{equation}
However, for $d=1,2,3$ this is convergent as $a \to 0$. Thus, recalling
\eq{Jscale}, we obtain
\begin{equation}
P_{\mathrm{split}} = O \left( \frac{J^2}{\tau^2} \right) =
\left\{ \begin{array}{ll} \tilde{O} \left( \frac{1}{\tau} \right) \,,
& d=1,2 \,,\\
\tilde{O} \left( \frac{a}{\tau} \right) \,, & d=3 \,.
\end{array} \right.
\end{equation}
We can consider two criteria regarding diabatic particle creation. If
our detectors are localized, we may be able to tolerate a low constant
density of stray particles created during state preparation. This
background is similar to that encountered in experiments, and may not
invalidate conclusions from the simulation. Alternatively, one could adopt a
strict criterion by demanding that, with high probability, not even
one stray particle is created in the volume being simulated during
state preparation. This strict criterion can be quantified by
demanding that the adiabatically produced state has an inner product of at
least $1-\epsilon$ with the exact state. This parameter $\epsilon$ is
thus directly comparable with that used in \sect{qubits}, and
the two sources of error can be added. Applying the strict criterion,
we demand that $P_{\mathrm{split}}$ and $P_{\mathrm{create}}$ each be
of order $\epsilon$, and obtain
\begin{equation}
\tau_{\mathrm{strict}} = \tilde{O} \left(
\frac{V}{\epsilon} \right) \,, \quad d=1,2,3 \,.
\end{equation}
Applying the more lenient criterion that $P_{\mathrm{create}}/V$ and
$P_{\mathrm{split}}$ each be of order $\epsilon$ yields
\begin{equation}
\tau_{\mathrm{lenient}} = \tilde{O} \left( \frac{1}{\epsilon} \right) \,,
\quad d=1,2,3 \,.
\end{equation}
For a $k\th$-order Suzuki-Trotter formula, the asymptotic scaling of the
total number of gates needed for adiabatic state preparation is
$O \left( (\mathcal{V} \tau)^{1+\frac{1}{2k}} \right) = O \left( (V
\tau/a^d)^{1+\frac{1}{2k}}) \right)$. Thus,
\begin{eqnarray}
\label{Gstrict}
G_{\mathrm{adiabatic}}^{\mathrm{strict}} & = & \tilde{O} \left(
\left( \frac{V^2}{a^d \epsilon} \right)^{1+\frac{1}{2k}} \right) \,, \\
\label{Glenient}
G_{\mathrm{adiabatic}}^{\mathrm{lenient}} & = & \tilde{O}
\left( \left( \frac{V}{a^d \epsilon} \right)^{1+\frac{1}{2k}} \right) \,.
\end{eqnarray}
\subsubsection{Strong Coupling}
\label{strong}
In two and three spacetime dimensions, we can obtain a strongly coupled
(that is, nonperturbative) field theory by approaching the
phase transition (\sect{QPT}). As in the case of weak
coupling, the necessary time for adiabatic state preparation depends
on various physical parameters of the system being simulated,
including the momentum of the incoming particles, the volume, the
strength of the final coupling, the number of spatial dimensions, and
the physical mass. To keep the discussion concise, we restrict our
discussion to the case of ultrarelativistic incoming particles, with
coupling strength close to the critical value. Under these conditions,
the incoming particles can produce a shower of many
$(n_{\mathrm{out}} \sim p/m)$ outgoing particles. Because of the strong
coupling, perturbation theory is inapplicable, and, even
if it could be used, would take exponential computation in the number of
outgoing particles.
In the strongly coupled case, we vary the Hamiltonian \ref{HS}
with $s$ by keeping the bare mass constant at $m_0$ and setting the
bare coupling to $s \lambda_0$. We choose $\lambda_0$ only slightly
below the critical value $\lambda_c$, so that at $s=1$ the system
closely approaches the phase transition, as illustrated in
Fig.~\ref{paths}. Examining \eq{thetap2} suggests that we can estimate
phase errors by understanding the behavior of $m^2(s)$ at $s=0$ and
$s=1$, without needing to know exactly what happens in between.
From \eq{eq:path},
\begin{equation}
\left. \frac{d m^2}{d s} \right|_{s=0} = \left\{ \begin{array}{ll}
\frac{\lambda_0}{8 \pi} \log \left( \frac{64}{m_0^2 a^2} \right) &
d=1 \,, \\
\frac{25.379}{16
\pi^2} \frac{\lambda_0}{a} &
d=2 \,,
\end{array} \right.
\end{equation}
and,
from \eq{numass} and \eq{nu},
\begin{equation}
\left. \frac{d m^2}{d s} \right|_{s=1} \sim \left\{
\begin{array}{ll}
-2 (\lambda_c - \lambda_0) &
d=1 \,, \\
-1.26 (\lambda_c - \lambda_0)^{0.26} &
d=2 \,.
\end{array} \right.
\end{equation}
Thus, \eq{phasecrit} yields
\begin{equation}
\label{unconcrete}
J = \tilde{O} \left( \sqrt{ \frac{\tau \lambda_0}{a^{d-1} p^2 \mathcal{D}}}
\right) \,, \quad d=1,2 \,,
\end{equation}
under the assumption that $(\lambda_c - \lambda_0)$ is very
small.
The result \ref{thetap2} rests on two approximations, a Taylor
expansion to second order in \eq{thetaj}, and an
approximation of a sum by an integral in \eq{thetap1}. The validity
conditions for these approximations become most stringent at $s=1$,
where the derivatives of $m^2$ with respect to $s$ become large.
Working out the $O(J^{-4})$ term in \eq{thetap2} at $s=1$, one
finds that it will be much smaller than the $O(J^{-2})$ term at $s=1$
provided
\begin{equation}
\label{Taylorcrit}
J \gg \frac{1}{\lambda_c - \lambda_0} \,.
\end{equation}
Similarly, higher-order terms in the Taylor expansion are suppressed
by additional powers of $\frac{1}{J(\lambda_c-\lambda_0)}$. The criterion
\ref{Taylorcrit} also suffices to justify the approximation of the sum
by an integral in \eq{thetap1}.
We must next consider adiabaticity to determine $\tau$. In
the ultrarelativistic limit, the relevant energy gap $\gamma$ is $\sim
\frac{m^2}{p}$. This takes its minimum value at $s=1$, namely,
\begin{equation}
\label{gammamin}
\gamma_{\min} \simeq \left\{ \begin{array}{ll}
\frac{(\lambda_c - \lambda_0)^2}{p} \,, & d=1 \,, \\
\frac{(\lambda_c - \lambda_0)^{1.26}}{p} \,, & d=2 \,.
\end{array} \right.
\end{equation}
Unlike in the perturbative case, we cannot make a detailed quantitative
analysis, but under the condition~\ref{Taylorcrit}, we should again be able
to apply the traditional adiabatic criterion and obtain a diabatic
transition amplitude scaling as $\frac{J}{\tau \gamma^2}$. Thus, to
keep the error probability at some small constant $\epsilon$, we have
\begin{equation}
\label{tauscale}
\tau \sim \frac{J}{\gamma^2 \sqrt{\epsilon}} \,.
\end{equation}
We now consider asymptotic scaling with $p$ for fixed
$\lambda_0$. To achieve continuum-like behavior we need $a \ll
\frac{1}{p}$. Thus \eq{unconcrete} yields
\begin{equation}
\label{pscaleJ}
J \sim \tau^{1/2} p^{(d-3)/2} \,, \quad d=1,2 \,.
\end{equation}
Substituting \eq{Taylorcrit} and \eq{gammamin} into
\eq{tauscale}, we see that we need
\begin{equation}
\label{cond1}
\tau \gtrsim p^2 \,, \quad d=1,2 \,.
\end{equation}
Substituting \eq{pscaleJ} and \eq{gammamin} into
\eq{tauscale}, we see that we need
\begin{equation}
\label{cond2}
\tau \gtrsim p^{d+1} \,, \quad d=1,2 \,.
\end{equation}
The scaling $\tau = O(p^{d+1})$ for $d=1,2$ suffices to satisfy both
conditions~\ref{cond1} and \ref{cond2}. Thus, by \S \ref{Trotter},
the total number of gates scales as
\begin{eqnarray}
G_{\mathrm{strong}} & = & O((V\tau)^{1+o(1)}p^{d+1+o(1)}) \\
& = & O \left( V^{1+o(1)} p^{2d+2+o(1)} \right) \,, \label{strongpscale}
\end{eqnarray}
for $d=1,2$.
Next, we consider asymptotic scaling with $(\lambda_c - \lambda_0)$
for fixed $p$. The $J$ scaling as $\sqrt{\tau}$ in
\eq{unconcrete} automatically satisfies the condition
\ref{Taylorcrit}. Thus, we substitute \eq{unconcrete} into
\eq{tauscale}, obtaining
\begin{equation}
\tau \sim \left\{ \begin{array}{ll} \left( \frac{1}{\lambda_c -
\lambda_0} \right)^8 \,, & d=1 \,, \\
\left( \frac{1}{\lambda_c - \lambda_0} \right)^{5.04} \,, & d=2 \,.
\end{array} \right.
\end{equation}
Thus, using a $k\th$-order Suzuki-Trotter formula, we obtain
\begin{equation}
\label{stronglambdascale}
G_{\mathrm{strong}} \sim \left\{ \begin{array}{ll} \left( \frac{1}{\lambda_c -
\lambda_0} \right)^{8 \left( 1+\frac{1}{2k} \right)} \,, & d=1 \,, \\
\left( \frac{1}{\lambda_c - \lambda_0} \right)^{5.04
\left(1+\frac{1}{2k} \right) } \,, & d=2 \,.
\end{array} \right.
\end{equation}
Note that one could improve this scaling by choosing a more optimized
adiabatic state-preparation schedule, which slows down as the gap gets
smaller.
\subsection{Suzuki-Trotter Formulae for Large Lattices}
\label{Trotter}
It appears that, while scaling with $t$ has been thoroughly studied,
little attention has been given to scaling of quantum simulation
algorithms with the number of lattice sites $\mathcal{V}$.
Using a result of Suzuki and elementary Lie algebra theory, we derive
linear scaling provided the Hamiltonian is local.
For any even $k$ and any pair of Hamiltonians $A, B$,
\begin{equation}
\label{hightrotter}
\left( e^{i A \alpha_1 t/n} e^{i B \beta_1 t/n} e^{i A \alpha_2 t/n}
e^{i \beta_2 B t/n} \ldots e^{i A \alpha_r t/n} \right)^n = e^{i (A + B) t}
+ O(t^{2k+1}/n^{2k}) \,,
\end{equation}
where $r = 1+5^{k/2-1}$ and
$\alpha_1,\ldots,\alpha_r,\beta_1,\ldots,\beta_{r-1}$ are specially
chosen coefficients such that $\sum_{j=1}^r \alpha_j = 1$ and
$\sum_{j=1}^{r-1} \beta_j = 1$ \cite{Suzuki}. Thus, using the $k\th$-order
Suzuki-Trotter formula (\eq{hightrotter}), one can simulate evolution
for time $t$ with $O\left( t^{\frac{2k+1}{2k}} \right)$ quantum gates
\cite{Cleve_sim}. To determine the $\mathcal{V}$ scaling, we use the
following standard theorem (cf. the Baker-Campbell-Hausdorff
formula).
\begin{theorem}\label{BCH}
Let $A$ and $B$ be elements of a Lie algebra defined over any field of
characteristic 0. Then $e^{A} e^{B} = e^{C}$, where $C$ is a formal
infinite sum of elements of the Lie algebra generated by $A$ and $B$.
\end{theorem}
$A$ and $B$ generate a Lie algebra by commutation and linear
combination. Thus, without requiring any explicit calculation,
Theorem~\ref{BCH} together with \eq{hightrotter} implies
\begin{equation}
\left( e^{i A \delta_1 t/n} e^{i B \delta_2 t/n} \ldots e^{i A
\delta_r t/n} \right)^n = e^{i (A + B) t} + \Delta_{2k+1}
t^{2k+1}/n^{2k} + O(n^{-(2k+1)}) \,,
\end{equation}
where $\Delta_{2k+1}$ is a linear combination of nested
commutators. In general, $\| \Delta_{2k+1} \|$ could be as large as
$\left( \max \left\{ \|A\|, \|B\| \right\} \right)^{2k+1}$. However,
by the canonical commutation relations, one sees that, for the pair of local
Hamiltonians $H_{\phi}, H_{\pi}$, $\|\Delta_{2k+1}\| =
O(\mathcal{V})$, for any fixed $k$. Thus, one needs only
$n = O \left( t^{\frac{2k+1}{2k}} \mathcal{V}^{\frac{1}{2k}}
\right)$. Recalling the $O(\mathcal{V})$ cost for simulating each
$e^{i H_{\phi} \delta t}$ or $e^{i H_{\pi} \delta t}$, one sees that
the total number of gates scales as $O\left( \left( t \mathcal{V}
\right)^{1+\frac{1}{2k}} \right)$. Note that this conclusion may be of
general interest, as it applies to any lattice Hamiltonian for which
non-neighboring terms commute.
In the case of strong coupling, we care not only about how the number
of gates scales with $\mathcal{V}$, but also about scaling with
$p$. In the presence of high-energy incoming particles, the field can
have large distortions from its vacuum state. For example, if
$\bra{\psi} \phi(\mathbf{x}) \ket{\psi}$ is large, then local terms in
$\Delta_{2k+1} \ket{\psi}$ such as $\pi(\mathbf{x}) \phi(\mathbf{x})^3
\ket{\psi}$ can become large. We can obtain a heuristic upper bound on
this effect by noting that, in the strongly coupled case, $m_0^2 > 0$,
so each local term in $H$ is a positive operator. Thus, if $\bra{\psi}
H \ket{\psi} \leq E$, then the expectation value of each of the local
terms is bounded above by $E$. Using $E$ as a simple estimate of the
maximum magnitude of a local term, we see that $\Delta_{2k+1}
\ket{\psi}$, which is a sum of $O(\mathcal{V})$ terms, each of which
is of degree $2k+1$ in the local terms of $H$, has magnitude at most
$O(\mathcal{V} E^{2k+1})$, or in other words $O(\mathcal{V}
p^{2k+1})$. Recalling that $a$ scales as a small multiple of $1/p$, we
see that $\Delta_{2k+1} \ket{\psi} = O(V p^{2k+1+d})$. Thus, $n =
O(p^{1+(1+d)/2k}t^{1+1/2k})$. Each timestep requires $O(\mathcal{V}) =
O(Vp^d)$ gates to implement. Thus, the overall scaling is
$O(p^{d+1+o(1)} (tV)^{1+o(1)})$ quantum gates to simulate the strongly
coupled theory at large $p$.
\end{document}
|
\section{Introduction}
The fact that the universe is expanding at every point in space is a
difficult concept to grasp. Cosmological observations from cosmic
microwave background radiation (CMBR) \cite{1} reveal that most of
the energy in our universe is dark which causes gravitational
repulsion and hence accelerates expansion of the universe. \cite{2}.
The properties of dark energy (DE) can be specified by energy
density $\rho$ and pressure $p$. These two parameters are
responsible for the following three main phases of the universe.
\begin{itemize}
\item The first phase is referred to as radiation dominated era
which occurred just after creation of the universe. At this stage
pressure of the radiation is given by one third of its energy
density.
\item The next is matter dominated era which came into being when universe
was assumed to be $70,000$ years old. Its energy density surpassed
the energy density of the first phase of the universe until a
cosmological constant as a DE candidate was proposed.
\item The third era is DE dominated era. About $5$ billion years ago this phase dominated
in the universe as a whole just after the matter dominated era which
was dropped to very low concentration \cite{3}. It is mentioned here
that the most recent Wilkinson Microwave Anisotropy Probe (WMAP)
observations \cite{2} indicate about $74\%$ of DE in the universe.
\end{itemize}
By the measurements of CMBR, the WMAP satellite indicates that the
universe is very close to the flat and to maintain this flatness,
the mass/energy density of the universe must be equal to a certain
critical density. The total amount of matter in the universe is
estimated only $30\%$ of the critical density. For the remaining
$70\%$ of critical density, an additional form of energy is
required which is termed as DE. There are two proposed forms to
discuss DE: one is the modified theories of gravity and the second
is the scalar field models. In this connection, single scalar
field models attracted many people. K-essence (k stands for
kinetic) model \cite{4} is one of the models which can be
described by a single real scalar field $\phi$ with non-canonical
kinetic term responsible for negative pressure. This may be taken
as the generalization of canonical scalar fields, e.g.,
quintessence \cite{5} (a time varying quantity).
The Lagrangian with a non-canonical kinetic term was proposed to
discuss the early time acceleration named as k-inflation \cite{6}.
Nojiri \cite{7} constructed explicitly k-essence DE model to unify
the late-time acceleration and inflation in early universe.
Matsumoto and Nojiri \cite{8} reconstructed the scalar quintessence
model, tachyon DE model, ghost condensation model as the special
cases of k-essence. Armendariz-Picon et al. \cite{9} discussed
essential features of k-essence and developed some examples. The
solution of these examples lead to two results: one in which cosmic
acceleration continues forever and the other in which acceleration
has finite duration. Bose and Majumdar \cite{10} investigated purely
kinetic k-essence and a particular k-essence model with potential
term. They concluded that such a model could generate basic features
of early inflation and DE observational constraints. Yang and Gao
\cite{11} introduced purely kinetic k-essence by means of
Lagrangian. They plotted evolutions of equation of state (EoS) and
speed of sound for particular cases.
Modified theories also play an important role in explaining DE.
There exist many modified gravity theories which may naturally unify
inflation in early universe and late-time acceleration \cite{12}.
$F(T)$ gravity \cite{13} is the modified form of teleparallel
equivalent of General Relativity \cite{14}. This theory is formed by
using Weitzenb$\ddot{o}$ck connection which has no curvature but
only torsion and possesses second order set of the field equations.
Myrzakulov \cite{15} proposed some new models of k-essence in the
framework of $F(T)$ gravity. Tsyba et al. \cite{16} investigated
purely kinetic k-essence for an explanation of cosmic
acceleration. They concluded that for a particular case of scalar
field, the modified gravity and purely kinetic k-essence become
equivalent. Dent et al. \cite{17} investigated this extended
modified gravity at the background and perturbed level and also
explored this theory for quintessence scenarios. Karami and
Abdolmaleki \cite{18} found that EoS parameter of holographic and
new agegraphic $F(T)$ models always cross the phantom-divide line
whereas entropy-corrected model has to experience some conditions
on parameters of the model. Most of the above mentioned work has
been carried out by using FRW universe for the obvious reasons.
In this paper, we discuss cosmic acceleration by using purely
kinetic k-essence and $F(T)$ gravity with Bianchi type $I$ universe.
The format of the paper is as follows: In the next section, we
present some basics of k-essence model and $F(T)$ gravity theory. In
section \textbf{3}, the field equations are formulated. Section
\textbf{4} is devoted to study some k-essence models and also
discuss cosmic acceleration. In the last section, some concluding
remarks are given.
\section{Preliminaries}
In this section, we provide basic concepts of k-essence as well as
$F(T)$ gravity.
\subsection{K-essence Formalism}
There are some scalar fields with non-canonical kinetic terms in
particle physics. The k-essence models are described by a single
scalar field and a kinetic term. The general k-essence action
\cite{6} is of the form
\begin{equation}\label{1}
S=\int d^{4}x\sqrt{-g}L(\phi,X),
\end{equation}
where $\phi$ is the scalar field, $g=det(g_{\mu\nu})$ and $X$ is the
dimensionless kinetic energy term defined by
\begin{equation}\label{2}
X=\frac{1}{2}g^{\mu\nu}\phi_\mu\phi_\nu,\quad \mu,\nu=0,1,2,3,
\end{equation}
where $\phi_\mu=\frac{\partial\phi}{\partial
x^\mu}=\partial_\mu\phi$. It is mentioned here that a Lagrangian can
be a function of any scalar field $\phi$ and $X$, i.e.,
$L=K(\phi,X)$. Here we consider one of the simplest possible
k-essence model, termed as, purely kinetic k-essence with action
\begin{equation}\label{3}
S=\int d^{4}x\sqrt{-g}L(X),\quad L=K(X).
\end{equation}
The energy-momentum tensor of k-essence is obtained by varying this
action with respect to the metric tensor \cite{6}
\begin{equation}\label{4}
T_{\mu\nu}=K_{X}\phi_{\mu}\phi_{\nu}-Kg_{\mu\nu},\quad
K_X=\frac{dK}{dX}.
\end{equation}
The energy-momentum tensor of perfect fluid is
\begin{equation}\label{5}
T_{\mu\nu}=(\rho_k+p_k)u_\mu u_\nu-p_{k}g_{\mu\nu}.
\end{equation}
Here, $\rho_k,~p_k$ are the energy density and pressure of k-essence
respectively and $u_\mu$ is the four-velocity defined by
\begin{equation}\label{6}
u_\mu=\eta\frac{\phi_\mu}{\sqrt{2X}},
\end{equation}
where $\eta=\pm1$ according to the sign of $\dot{\phi}$ positive or
negative respectively. We assume that $\partial_{\mu}\phi$ is
timelike and smooth on interesting scales \cite{5}. Thus we can
associate the energy-momentum tensors of k-essence and perfect
fluid. For $\mu=0=\nu$ and $\mu=1=\nu$ in Eqs.(\ref{4}) and
(\ref{5}), we obtain the following expressions for k-essence energy
density and pressure respectively
\begin{eqnarray}\label{8}
\rho_{k}=2XK_{X}-K,\quad p_k=K.
\end{eqnarray}
This yields the following k-essence EoS parameter
\begin{equation}\label{10}
\omega_k=\frac{p_k}{\rho_k}=\frac{K}{2XK_{X}-K}.
\end{equation}
\subsection{$F(T)$ Theory of Gravity}
Here, we introduce briefly the teleparallel theory of gravity and
its generalization to $F(T)$ gravity. In teleparallel action, the
torsion scalar $T$ is used as the Lagrangian density while in
modified teleparallel gravity, it is promoted to a function of $T$.
Thus the action for $F(T)$ gravity \cite{18} is
\begin{equation}\label{11}
S=\frac{1}{2\kappa^2}\int d^{4}x[eF(T)+L_m],
\end{equation}
where $e=\sqrt{-g},~\kappa^{2}=8\pi G,~G$ is the gravitational
constant and $F(T)$ is the general differentiable function of $T$.
Also, $L_m$ is the Lagrangian density of matter inside the
universe. The torsion scalar is given as
\begin{equation}\label{12}
T=S_{\rho}~^{\mu\nu}T^{\rho}~_{\mu\nu},
\end{equation}
where $S_{\rho}~^{\mu\nu}$ and torsion tensor $T^{\rho}~_{\mu\nu}$
are defined as follows
\begin{eqnarray}\label{13}
S_{\rho}~^{\mu\nu}&=&\frac{1}{2}(K^{\mu\nu}~_{\rho}
+\delta^{\mu}_{\rho}T^{\theta\nu}~_{\theta}-\delta^{\nu}_{\rho}T^{\theta\mu}~_{\theta}),\\
\label{14}T^{\lambda}~_{\mu\nu}&=&\Gamma^{\lambda}~_{\nu\mu}-
\Gamma^{\lambda}~_{\mu\nu}=h^{\lambda}_{i}
(\partial_{\mu}h^{i}_{\nu}-\partial_{\nu}h^{i}_{\mu}).
\end{eqnarray}
Here $h^{i}_{\mu}$ are tetrad components which form an orthonormal
basis for the tangent space at each point $x^\mu$ of the manifold.
Each vector $h_i$ can be identified by its components $h_\mu^i$
such that $h_i=h^\mu_i \partial_\mu$, where index $i$ runs over
$0,1,2,3$ denote the tangent space while $\mu=0,1,2,3$ represent
the coordinate indices on the manifold. These tetrad related to
the metric tensor $g_{\mu\nu}$ by the following relation
\begin{equation}\label{4*}
g_{\mu\nu}=\eta_{ij}h_{\mu}^{i}h_{\nu}^{j},
\end{equation}
where $\eta_{ij}=diag(1,-1,-1,-1)$ is the Minkowski metric for the
tangent space satisfying the following properties
\begin{equation}\label{11*}
h^{i}_{\mu}h^{\mu}_{j}=\delta^{i}_{j},\quad
h^{i}_{\mu}h^{\nu}_{i}=\delta^{\nu}_{\mu}.
\end{equation}
The contorsion tensor, $K^{\mu\nu}_{\rho}$ is equal to the
difference between Weitzenb$\ddot{o}$ck and Levi-Civita connection
defined by
\begin{equation}\label{15}
K^{\mu\nu}~_{\rho}=-\frac{1}{2}(T^{\mu\nu}~_{\rho}
-T^{\nu\mu}~_{\rho}-T_{\rho}~^{\mu\nu}).
\end{equation}
The variation of Eq.(\ref{11}) with respect to tetrad $h_\mu^i$
leads to the following field equations
\begin{equation}\label{16}
[e^{-1}\partial_{\mu}(eS_{i}~^{\mu\nu})-
h^{\lambda}_{i}T^{\rho}~_{\mu\lambda}S_{\rho}~^{\nu\mu}]F_{T}
+S_{i}~^{\mu\nu}\partial_{\mu}(T)
F_{TT}+\frac{1}{4}h^{\nu}_{i}F=\frac{1}{2}\kappa^{2}h^{\rho}_{i}T^{\nu}_{\rho},
\end{equation}
where $F_{T}=\frac{dF}{dT},~F_{TT}=\frac{d^{2}F}{dT^{2}},~
S_{i}~^{\mu\nu}=h^{\rho}_{i}S_{\rho}~^{\mu\nu}$ with antisymmetric
property. The energy-momentum tensor $T_{\mu\nu}$ is given as
\begin{equation}\label{17}
T^\mu_\nu=diag(\rho_m,-p_m,-p_m,-p_m),
\end{equation}
where $\rho_{m}$ and $p_{m}$ denote the usual density and pressure
of matter inside the universe.
\section{Bianchi $I$ Universe and the Field Equations}
The assumption of isotropy does not predict early epoch of the big
bang as the universe does not maintain its isotropic behavior at
very small scales. In order to get a realistic model which
represents an expanding, homogenous and anisotropic universe, we use
Bianchi type I universe model which is a generalization of FRW
metric. Here we study evolution of the universe in the presence of
DE k-essence models. The line element of Bianchi type I spacetime
is given by
\begin{equation}\label{18}
ds^{2}=dt^{2}-A^{2}(t)dx^{2}-B^{2}(t)dy^{2}-C^{2}(t)dz^{2},
\end{equation}
where the scale factors $A,~B$ and $C$ are functions of cosmic
time $t$ only. Using Eqs.(\ref{4*}) and (\ref{18}), we obtain the
tetrad components as follows \cite{18a}
\begin{equation}\label{19}
h^{i}_{\mu}=diag(1,A,B,C).
\end{equation}
Using Eqs.(\ref{13}) and (\ref{14}) in (\ref{12}) along with
(\ref{18}), we get
\begin{equation}\label{20}
T=-2\left(\frac{\dot{A}\dot{B}}{AB}+\frac{\dot{B}\dot{C}}{BC}+\frac{\dot{C}\dot{A}}{CA}\right).
\end{equation}
The field equations (\ref{16}) for $i=0=\nu$ and $i=1=\nu$ are given
by
\begin{eqnarray}\label{21}
F-4\left(\frac{\dot{A}\dot{B}}{AB}+\frac{\dot{B}\dot{C}}{BC}
+\frac{\dot{C}\dot{A}}{CA}\right)F_{T}=2\kappa^{2}\rho_{m},
\end{eqnarray}
\begin{eqnarray}\nonumber
&&2\left(\frac{\dot{A}\dot{B}}{AB}+2\frac{\dot{B}\dot{C}}{BC}
+\frac{\dot{C}\dot{A}}{CA}+\frac{\ddot{B}}{B}+\frac{\ddot{C}}{C}\right)F_{T}
-4\left(\frac{\dot{B}}{B}+\frac{\dot{C}}{C}\right)\\\nonumber
&&\times\left[\left(\frac{\ddot{A}}{A}
-\frac{\dot{A^{2}}}{A^{2}}\right)\left(\frac{\dot{B}}{B}+\frac{\dot{C}}{C}\right)
+\left(\frac{\ddot{B}}{B}-\frac{\dot{B^{2}}}{B^{2}}\right)\left(\frac{\dot{C}}{C}
+\frac{\dot{A}}{A}\right)\right.\\\label{22}
&&\left.+\left(\frac{\ddot{C}}{C}-\frac{\dot{C^{2}}}{C^{2}}\right)
\left(\frac{\dot{A}}{A}+\frac{\dot{B}}{B}\right)\right]F_{TT}-F=2\kappa^{2}p_{m}.
\end{eqnarray}
The corresponding conservation equation is
\begin{equation}\label{23}
\dot{\rho_{m}}+\left(\frac{\dot{A}}{A}+\frac{\dot{B}}{B}+\frac{\dot{C}}{C}\right)
(\rho_{m}+p_{m})=0.
\end{equation}
The average scale factor $R$ and the mean Hubble parameter $H$
respectively will become
\begin{eqnarray}\label{24}
R=(ABC)^{1/3},\quad
H=\frac{1}{3}\left(\frac{\dot{A}}{A}+\frac{\dot{B}}{B}+\frac{\dot{C}}{C}\right)
=\frac{\dot{R}}{R}.
\end{eqnarray}
The deceleration parameter $q$ is a dimensionless quantity and is
used to describe the accelerating universe
\begin{equation}\label{q}
q=-1-\frac{\dot{H}}{H^2}.
\end{equation}
Any cosmological universe represents an accelerating, decelerating
or expansion with constant velocity for $-1\le q<0$, $q>0$ and $q=0$
respectively. In terms of Hubble parameter and torsion,
Eqs.(\ref{21})-(\ref{23}) can be simplified as
\begin{eqnarray}\label{26}
2\kappa^{2}\rho_m&=&2TF_T+F,\\\label{27}
2\kappa^2p_m&=&(6\dot{H}-T+2J+2L)F_T+2M\dot{T}F_{TT}-F,\\\label{28}
0&=&\dot{\rho}_m+3H(\rho_m+p_m),
\end{eqnarray}
where $L=\frac{\dot{B}\dot{C}}{BC}-\frac{\ddot{A}}{A},~
M=\frac{\dot{B}}{B}+\frac{\dot{C}}{C}$ and
$J=\frac{\dot{A^{2}}}{A^2}+\frac{\dot{B^{2}}}{B^2}+\frac{\dot{C^{2}}}{C^2}$.
Also, Eq.(\ref{20}) can be written as
\begin{equation}\label{29}
T=-9H^{2}+J.
\end{equation}
The case $F(T)=T$ gives the torsion contribution $\rho_{T},~p_{T}$
in Eqs.(\ref{26}) and (\ref{27}) which reduce to the following forms
\begin{eqnarray}\label{30}
&&\rho_{m}+\rho_{T}=\frac{3T}{2\kappa^{2}},\\\label{31}
&&p_{m}+p_{T}=\frac{1}{\kappa^2}\left(3\dot{H}-T+J+L\right).
\end{eqnarray}
If we take $\rho_T=\rho_k,~p_T=p_k$, then these equations become
\begin{eqnarray}\label{33}
&&\rho_{m}+2XK_{X}-K=\frac{3T}{2\kappa^{2}},\\\label{34}
&&p_{m}+K=\frac{1}{\kappa^2}\left(3\dot{H}-T+J+L\right).
\end{eqnarray}
For the sake of simplicity, we assume that
$\rho_m=0=p_m=\kappa^2-1$. Using these values and inserting the
value of $T$ from Eq.(\ref{29}) in (\ref{33}) and (\ref{34}), it
follows that
\begin{eqnarray}\label{37}
2XK_{X}-K&=&\frac{3}{2}\left(-9H^2+J\right),\\\label{38}K
&=&3\dot{H}+9H^2+L.
\end{eqnarray}
Equation (\ref{2}) implies that $X=\frac{1}{2}\dot{\phi}^2$ whose
corresponding scalar function is given by
\begin{equation}\label{7}
\phi=\int{\sqrt{2X}}dt+constant.
\end{equation}
The continuity equation (\ref{5}) for k-essence models leads to
\begin{equation}\label{39}
\dot{\rho_{k}}+3H(\rho_{k}+p_{k})=0.
\end{equation}
Using Eq.(\ref{8}) in this equation, we have
\begin{equation}\label{40}
(K_X+2XK_{XX})\dot{X}+6HXK_X=0
\end{equation}
which is the kinetic k-essence field equation.
As a special case \cite{15}, we take $\phi=\phi_0+\ln
R^{\pm\sqrt{18}}$, where $\phi_0$ is an arbitrary constant.
Consequently, the kinetic term takes the form
\begin{equation}\label{40+}
X=\frac{1}{2}\dot{\phi}^2=J-T.
\end{equation}
Inserting this value in Eq.(\ref{40}), we get
\begin{equation}\label{40++}
2T(\dot{J}-18H\dot{H})F_{TT}+(\dot{J}-18H\dot{H}+6HT)F_T=0.
\end{equation}
Also, Eq.(\ref{27}) can be simplified as
\begin{equation}\label{28+}
2M(\dot{J}-18H\dot{H})F_{TT}+(6\dot{H}+9H^2+J+2L)F_T-F=0.
\end{equation}
It is interesting to mention here that if we use Eq.(\ref{40++}) in
(\ref{28+}), it shows the equivalence of $F(T)$ gravity and
k-essence for FRW spacetime \cite{15},\cite{16}. However, these
equation do not provide any such relation for Bianchi type $I$
universe.
\section{K-essence Models}
In this section, we aim to discuss k-essence models in the framework
of modified teleparallel gravity. We take three arbitrary k-essence
models and evaluate kinetic term and the corresponding scalar field.
Also, we formulate EoS parameter and the deceleration parameter for
k-essence models. For this purpose, we take the following particular
values of the scale factors by using the power law \cite{20}
\begin{equation}\label{52}
A(t)=(m_{1}s_{1}t)^{1/m_1},\quad B(t)=(m_{2}s_{2}t)^{1/m_2},\quad
C(t)=(m_{3}s_{3}t)^{1/m_3},
\end{equation}
where $m_i,~s_i$ are positive constants and $i=1,2,3$. The EoS
parameter in Eq.(\ref{10}) can be written as follows
\begin{equation}\label{50}
\omega_k=-1+\frac{2XK_X}{2XK_X-K}.
\end{equation}
\subsection{Model I}
Consider the following k-essence model \cite{16}
\begin{equation}\label{41}
K=\sum^{M}_{j=0}\nu_j(t)y^j,\quad M>0,
\end{equation}
where $y=\tanh{t}$. We find the coefficients of $y$ and the kinetic
term of k-essence ($X$) by assuming the following Hubble parameter
\begin{equation}\label{42}
H=\sum_{j=0}^{N}\mu_j(t)y^j,\quad N>0.
\end{equation}
Notice that we take $\mu_j(t)$ and $\nu_j(t)$ as constant
quantities throughout. As a particular example, we take $M=2$ in
Eq.(\ref{41}) and $M=1$ for $H$, it follows that
\begin{eqnarray}\label{43}
K=\nu_0+\nu_{1}y+\nu_{2}y^2,\quad H=\mu_0+\mu_{1}y.
\end{eqnarray}
Inserting this value of $H$ in Eq.(\ref{38}), we obtain
\begin{equation}\label{45}
K=3\mu_{1}(3\mu_{1}-1)y^{2}+18\mu_{0}\mu_{1}y+3\mu_{1}+9\mu_{0}^2+L.
\end{equation}
Using the values of $K$ and $H$ in Eq.(\ref{37}), it follows that
\begin{equation}\label{46}
X=a_{1}e^{\int g_1(t)dt},
\end{equation}
where $a_1$ is an integration constant and
\begin{eqnarray}\nonumber
g_1(t)&=&[3\mu_1+\frac{45}{2}\mu_0^2+L-\frac{3}{2}J-9\mu_0\mu_1y-3\mu_1
(\frac{3}{2}\mu_1+1)y^2]^{-1}\\\nonumber
&\times&2[18\mu_0\mu_1+3\mu_1+9\mu_0^2+\dot{L}+6\mu_1
(3\mu_1-1)y-18\mu_0\mu_1y^2\\\label{47}
&-&6\mu_1(3\mu_1-1)y^3].
\end{eqnarray}
It is mentioned here that the scalar function $\phi$ for the model
(\ref{41}) can be obtained by using Eqs.(\ref{7}) and (\ref{46}).
The k-essence energy density and pressure in Eq.(\ref{8}) take the
following forms
\begin{eqnarray}\label{48}
&&\rho_k=-\frac{27}{2}\mu_1^2y^2-27\mu_0\mu_1y-\frac{27}{2}\mu_1^2+\frac{3}{2}J,\\\label{49}
&&p_k=3\mu_{1}(3\mu_{1}-1)y^{2}+18\mu_{0}\mu_{1}y+3\mu_{1}+9\mu_{0}^2+L.
\end{eqnarray}
Inserting these values in Eq.(\ref{50}), the EoS parameter becomes
\begin{equation}\label{51}
\omega_k=-1-\frac{3\mu_1-\frac{9}{2}\mu_0^2+L+\frac{3}{2}J-9\mu_0\mu_1y-3\mu_1
(\frac{3}{2}\mu_1+1)y^2}{\frac{3}{2}(9\mu_{0}^2-J)+27\mu_0\mu_1y+\frac{27}{2}\mu_1^2y^2}.
\end{equation}
The viability of this model depends on the possible values of the
parameters in Eq.(\ref{52}). The model shows different behavior
initially for unequal $m_{i}$ but indicates the same behavior at
later times for all values of the parameters. Here we discuss the
simple case which leads to the isotropic universe. Using
Eq.(\ref{52}) in this equation and assuming $\mu_i=1=m_i=s_i$, it
follows that
\begin{equation}\label{53}
\omega_k=-1+\frac{1-11/3t^2+6\tanh{t}+5\tanh^2{t}}{9-3/t^2+18\tanh{t}+9\tanh^2{t}}.
\end{equation}
For late time acceleration, i.e., $t\rightarrow \infty$, this yields
\begin{equation}\label{54}
\omega_k\mid_{t\rightarrow\infty}=-\frac{2}{3}
\end{equation}
\begin{figure}
\center\epsfig{file=f1.eps, width=0.5\linewidth}\caption{Plot of
$\omega_k$ versus cosmic time $t$ for model I.}
\center\epsfig{file=f11.eps, width=0.5\linewidth}\caption{Plot of
$q$ versus cosmic time $t$ for model I.}
\end{figure}
which is shown in \textbf{Figure \ref{1}}. This represents evolution
of k-essence EoS parameter $\omega_k$ as a function of cosmic time.
It is obvious that initially $(t=0),~\omega_k=0.22$ showing that the
universe is lying in a region which contained dust fluid as well as
radiation. As the time elapses up to $t=0.42$, k-essence EoS
parameter confined in the region $-\infty<\omega_k<+\infty$, i.e.,
the universe evolutes from physical matter to DE phase. The EoS
parameter bears a negative increment in its value and becomes
constant at $t=0.48$ towards $t\rightarrow\infty$. The constant
value of EoS parameter is greater than $-1$ but less than $-1/3$
which shows a quintessence era \cite{21}.
The deceleration parameter (\ref{q}) for this model takes the form
\begin{equation}\label{q1}
q=-1+\frac{\mu_1(1-\tanh^2{t})}{\mu_0^2+\mu_1^2\tanh^2{t}+2\mu_0\mu_1\tanh{t}}
\end{equation}
which implies that $q\mid_{t\rightarrow\infty}=-1$ indicating an
accelerating universe. Its graphical representation is shown in
\textbf{Figure \ref{2}} which indicates that initially at
$(t=0),~q=0$ expressing an expanding universe with a constant
velocity. As time passes, its value decreases and converges towards
$-1$. Notice that the graph exhibits an ever expanding universe as
there is not a single positive value of $q$.
\subsection{Model II}
The second k-essence model \cite{15} is given by
\begin{equation}\label{55}
K=\sum^{n}_{j=-m}\nu_{j}(t)e^{jt},
\end{equation}
where $m$ is any positive real number and $n\leq0$ real number. For
the sake of simplicity, we take $m=2,~n=0$ and $\nu_j(t)=\nu_j$ as a
constant. Thus
\begin{equation}\label{56}
K=\nu_{-2}e^{-2t}+\nu_{-1}e^{-t}+\nu_0.
\end{equation}
We also assume $H$ as follows
\begin{equation}\label{57}
H=\mu_{-1}e^{-t}+\mu_0.
\end{equation}
Inserting this value in Eq.(\ref{38}), we get
\begin{equation}\label{58}
K=9\mu_0^2+L-3\mu_{-1}(1-6\mu_0)e^{-t}+9\mu_{-1}^2e^{-2t}.
\end{equation}
Using Eqs.(\ref{57}) and (\ref{58}) in (\ref{37}), we obtain kinetic
term of k-essence model in the form
\begin{equation}\label{59}
X=a_2e^{\int g_2(t)dt}.
\end{equation}
Here $a_2$ is an integration constant and $g_2(t)$ is given by
\begin{equation}\label{60}
g_2(t)=\frac{2[\dot{L}+3\mu_{-1}(1-6\mu_0)e^{-t}-18\mu_{-1}^2e^{-2t}]}{3J/2-9/2+L-3\mu_{-1}
(1+3\mu_{0})e^{-t}-9\mu_{-1}^2e^{-2t}}.
\end{equation}
The scalar function $\phi$ is obtained by inserting Eq.(\ref{59}) in
(\ref{7}).
\begin{figure}
\center\epsfig{file=f2.eps, width=0.5\linewidth}\caption{Plot of
$\omega_k$ versus cosmic time for model II.}
\center\epsfig{file=f22.eps, width=0.5\linewidth}\caption{Plot of
$q$ versus cosmic time for model II.}
\end{figure}
The energy density and pressure parameters turn out to be
\begin{eqnarray}\label{61}
\rho_k&=&-\frac{27}{2}\mu_0^2+\frac{3}{2}J-27\mu_{-1}\mu_0e^{-t}-\frac{27}{2}
\mu_{-1}^2e^{-2t},\\\label{62}p_k&=&9\mu_0^2+L-3\mu_{-1}(1-6\mu_0)e^{-t}+9\mu_{-1}^2e^{-2t}.
\end{eqnarray}
Inserting these values in Eq.(\ref{50}), the EoS parameter takes the
form
\begin{equation}\label{63}
\omega_k=-1+\frac{9\mu_0^2-2L-3J+6\mu_{-1}(1+3\mu_{0})e^{-t}+9\mu_{-1}^2e^{-2t}}
{27\mu_0^2-3J+54\mu_{-1}\mu_{0}e^{-t}+27\mu_{-1}e^{-2t}}.
\end{equation}
Taking all the constants equal to $1$ as in isotropic case and
using Eq.(\ref{52}), it follows that
\begin{equation}\label{64}
\omega_k=-1+\frac{9-11/t^2+24e^{-t}+9e^{-2t}}
{27-9/t^2+54e^{-t}+27e^{-2t}}.
\end{equation}
This shows that as $t\rightarrow\infty,~\omega_k>-1$ which
represents quintessence region \cite{21} shown in \textbf{Figure
\ref{3}}. This model has the same behavior as that of the first
model. However, the EoS parameter is directed to DE era at $t=0.34$
and after a short interval of time, it becomes constant, i.e.,
$-0.67$. The corresponding deceleration parameter is
\begin{equation}\label{q2}
q=-1+\frac{\mu_{-1}e^{-t}}{\mu_{-1}^2e^{-2t}+\mu_{0}^2+2\mu_{-1}\mu_{0}e^{-t}},
\end{equation}
which gives $q\mid_{t\rightarrow\infty}=-1$. At $t=0$, the value of
$q$ is negative which shows initially an expanding universe. Its
value decreases and converges to $-1$ as $t\rightarrow\infty$ with
the passage of time as shown in \textbf{Figure \ref{4}}.
\subsection{Model III}
Here we take the following k-essence model \cite{15}
\begin{equation}\label{65}
K=\sum^{n}_{j=-m}\nu_{j}(t)(\ln{t})^j,
\end{equation}
where $m,~n$ and $\nu_j$ are the same as in model II. For
$m=2,~n=0$ and constant $\nu_j$'s, this equation becomes
\begin{equation}\label{66}
K=\nu_{-2}(\ln{t})^{-2}+\nu_{-1}(\ln{t})^{-1}+\nu_0.
\end{equation}
Assuming the Hubble parameter in the form
\begin{equation}\label{67}
H=\mu_{-1}(\ln{t})^{-1}+\mu_0.
\end{equation}
Inserting this value of $H$ in Eq.(\ref{38}), it implies that
\begin{equation}\label{68}
K=L+9\mu_0^2+18\mu_{-1}\mu_0(\ln{t})^{-1}+3\mu_{-1}(3\mu_{-1}-1/t)(\ln{t})^{-2}.
\end{equation}
The kinetic term $X$ from (\ref{37}) and the corresponding scalar
function $\phi$ become
\begin{eqnarray}\label{69}
X=a_3e^{\int g_3(t)dt},\quad \phi=\partial_{t}^{-1}\sqrt{-2X},
\end{eqnarray}
where $g_3(t)$ is found by using (\ref{37})
\begin{eqnarray}\label{71}
&&g_3(t)=2[\dot{L}-6\mu_{-1}(3\mu_{-1}-1/t)t^{-1}(\ln{t})^{-3}+
3\mu_{-1}(1/t-6\mu_{0})t^{-1}(\ln{t})^{-2}]\nonumber\\
&&\times[L+3J/2-9\mu_0^2/2-9\mu_{-1}\mu_{0}
(\ln{t})^{-1}+3\mu_{-1}(-3\mu_{-1}/2-1/t)(\ln{t})^{-2}]^{-1}.\nonumber\\
\end{eqnarray}
\begin{figure} \center\epsfig{file=f3.eps,
width=0.5\linewidth}\caption{Plot of $\omega_k$ versus cosmic time
for model III.} \center\epsfig{file=f33.eps,
width=0.5\linewidth}\caption{Plot of $q$ versus cosmic time for
model III.}
\end{figure}
For this model, $\rho_k$ and $p_k$ become
\begin{eqnarray}\label{72}
\rho_k&=&\frac{3}{2}J-\frac{27}{2}\mu_0^2-27\mu_{-1}\mu_{0}(\ln{t})^{-1}-\frac{27}{2}
\mu_{-1}^2(\ln{t})^{-2},\\\label{73}
p_k&=&L+9\mu_0^2+18\mu_0\mu_{-1}(\ln{t})^{-1}+3\mu_{-1}(3\mu_{-1}-\frac{1}{t})(\ln{t})^{-2}.
\end{eqnarray}
The corresponding EoS parameter takes the form
\begin{equation}\label{74}
\omega_k=-1+\frac{2L+3J-9\mu_0^2-18\mu_{-1}\mu_0(\ln{t})^{-1}-6\mu_{-1}(3\mu_
{-1}/2+1/t)(\ln{t})^{-2}}{3J-27\mu_0^2-54\mu_{-1}\mu_0(\ln{t})^{-1}-27\mu_{-1}^2(\ln{t})^{-2}}.
\end{equation}
To expose the present day nature of the universe, we take
$\mu_i=m_i=s_i=1$ and using Eq.(\ref{52}), it follows that
\begin{equation}\label{75}
\omega_k=-1+\frac{3-11/3t^2+6(\ln{t})^{-1}+2(3/2+1/t)(\ln{t})^{-2}}
{9-3/t^2+18(\ln{t})^{-1}+9(\ln{t})^{-2}}.
\end{equation}
Consequently, the deceleration parameter become
\begin{equation}\label{q3}
q=-1+\frac{\mu_{-1}}{t(\ln{t})^2[\mu_{-1}^2(\ln{t})^{-2}+\mu_0^2+2\mu_0\mu_{-1}(\ln{t})^{-1}]}.
\end{equation}
This shows that the condition for accelerating universe is satisfied
as $t\rightarrow\infty$. \textbf{Figure \ref{5}} (\ref{75}) shows
that $\omega_k>-1$ as $t\rightarrow\infty$ and initially the model
represents a universe having both properties of matter and
radiation. After a very short interval, universe is dominated by DE.
It is worthwhile to mention here that at $t=0.63$, the universe
changed its phase from DE era to a physical matter dominated era. As
compared to the DE phase, the universe stayed in matter dominated
era for a short interval of time. Then a decrement in its value is
observed, the EoS parameter for k-essence attains negative values
and finally converges to a constant value resulting quintessence
region \cite{22} of the universe. \textbf{Figure \ref{6}} (\ref{q3})
represents a decelerating universe at early times as the value of
$q$ is positively increasing or decreasing at that stage. After a
long interval of time, the universe intersects the boundary from
matter to DE era and takes $q=-1$. This corresponds to an
accelerating expanding universe.
\section{Summary}
This paper is devoted to study the well-known phenomenon of the
universe expansion in the context of $F(T)$ gravity. For this
purpose, we have taken Bianchi type $I$ universe and have explored
some purely kinetic k-essence models. In physical cosmology and
gravity, the DE problem has been investigated in terms of
cosmological constant, scalar fields, tachyon, Chaplygin gas,
quintessence and modified gravities etc. Here have used $F(T)$
gravity and scalar field term whose combination has resulted a
quintessence phase.
We have evaluated the EoS parameter and the deceleration parameter
for purely kinetic k-essence models. These are shown in terms of
graphs versus cosmic time by taking particular values of the metric
coefficients. These models describe evolution of the universe from
big-bang to present epoch and are summarized as follows
\begin{itemize}
\item For model I, the EoS parameter and the deceleration parameter indicate an
ever accelerating universe with the passage of time. At $t=0.42$,
the universe changes its phase from matter dominated to DE phase
resulting quintessence region.
\item The model II has the same behavior as that of the model I, however,
phase changes at point $t=0.38$. The deceleration parameter
initially represents an expanding universe with constant speed and
after a short interval, it accelerates with a faster speed.
\item For model III, $\omega_k$ and $q$ initially show decelerating
DE dominated universe which reverses to matter dominated
decelerating universe at $t=0.64$. After a slight increment in time,
the universe shows accelerating recent epoch of DE issue lying in
the quintessence region.
\end{itemize}
The analysis of models reveal that the present day universe is
dominated by DE component which can successfully describe
accelerating nature of the universe consistent with the observations
\cite{23}. These new forms of models may give the equivalent
descriptions of DE to discuss the acceleration of the expanding
universe. These models indicate that the present day universe is
dominated by DE component. The viability of these models depend on
the possible values of parameters. Similarly, we can construct the
observational parameters for other new k-essence models induced by
modified gravity theory. Finally, it is worthwhile to mention here
that all the above results turn out to be the generalization of the
already obtained results for the FRW metric \cite{16}. The graphs of
models depend upon the values of parameters accordingly. For Bianchi
type $I$, the k-essence models in the context of generalized
teleparallel gravity mark out a quintessence DE phase.
|
\section{Methods}
\subsection{Atomistic Force Field}
Our atomistic simulations of the dominant folding trajectories of the Fip35 WW domain were performed using the AMBER~ff99SB force field~\cite{AMBER99sb} in implicit solvent with
Generalized Born formalism implemented in GROMACS 4.5.2 \cite{GRO4}. The Born radii were calculated according to the Onufriev-Bashford-Case algorithm \cite{OBC}.
In a recent work based on the DRP method, the dominant pathway in the conformational transition of tetra-alanine obtained using the same version of the AMBER
force field was found to agree well with the results of an analogous calculation in which the
molecular potential energy was determined {\it ab initio}, i.e. directly from quantum electronic structure calculations~\cite{QDRP2}.
\subsection{Coarse-grained Model}
To study the equilibrium properties of the folding of the Fip35 WW domain we used the coarse-grained model recently developed in Ref.s~\cite{Kim&Hummer,
reactioncoord2}. In that model,
aminoacids are represented by spherical beads centered at the $C_\alpha$ positions. The non-bonded part of the potential energy contains both native and non-native interactions.
The former are the same used in the G$\overline{o}$-type model of Ref.~\cite{Karanikolas&Brooks}, while the latter consist of a quasi-chemical potential,
which accounts for the statistical propensity of different aminoacids to be found in contact in native structures, and of a Debye-screened electrostatic term.
In this model, the
average potential energy due to native interactions in the folded phase is typically one order of magnitude larger than that due to non-native interactions.
Above the folding temperature, this ratio drops to about 4.
This model was shown to provide an accurate description of protein-protein complexes with low and intermediate binding affinities \cite{Kim&Hummer}. In the
insert of the first panel of Fig. \ref{Fig3} we plot the specific heat, evaluated from MC simulations at different temperatures, which indicates that this model
yields the correct folding temperature for this WW domain.
\subsection{The Dominant Reaction Pathways Method}
The high computational cost of MD simulations of macromolecular systems has triggered efforts towards developing alternative
theoretical frameworks to investigate their long-time dynamics and reaction kinetics (see e.g. \cite{DRP1, method1, method2, method3, method4, method5, method6} and references therein).
In particular, the DRP approach \cite{DRP1, DRP2, DRP3,DRP4, QDRP1, DRP0} concerns physical systems which can be described by the overdamped Langevin equation. If ${\bf x}_k$ denotes the
coordinate of the $k-$th atom, the Langevin equation in the so-called Ito Calculus reads:
\begin{equation} \label{langev_eq}
{\bf x}_k(i+1)= {\bf x}_k(i) -\frac{\Delta t D_k}{k_B T} {\bf \nabla}U[{\bf X}(i)] + \sqrt{2 D_k \Delta t}\, {\bf \eta}_k(i).
\end{equation}
In this equation, ${\bf X}(i)\equiv ({\bf x}_1(i),\ldots, {\bf x}_N(i))$ is the set of atomic coordinates at the $i-$th time step, $\Delta t$ is an elementary time interval, $D_k$ is the diffusion
coefficient of the $k-$th atom, $k_B$ is the Boltzmann's constant, $T$ is the temperature of the heat-bath,
$U({\bf X})$ is the potential energy. ${\bf \eta}^k(i)$ is a white Gaussian noise with unitary variance, acting on the $k-$th atom.
The probability for a protein to fold in a given time interval $t$ can be written as
\begin{eqnarray}
\label{Pf}
P_f(t) = \int d{\bf X}_f~h_N({\bf X}_f) \int d {\bf X}_i~h_D({\bf X}_i)~P({\bf X}_f, t| {\bf X}_i)~\rho_0({\bf X}_i),
\end{eqnarray}
where $h_{N(D)}({\bf X})$ is the characteristic function of the native (denatured) state (defined in terms of some suitable order parameters), $\rho_0({\bf X}_i)$ is the initial distribution of
micro-states in the denatured state and $P({\bf X}_f, t|{\bf X}_i)$ is the conditional probability of reaching the (native)
configuration ${\bf X}_f$ starting from the (denatured) configuration ${\bf X}_i$, in a time $t$. If the total time interval $t$ is chosen much smaller than the inverse folding rate,
this probability is dominated by single nonequilibrium folding events.
It can be shown that the probability of a given folding trajectory ${\bf X}(t)$ connecting denatured and native configurations is proportional to the negative exponent of the Onsager-Machlup functional \cite{DRP0, method5}, which in discretized form reads
\begin{eqnarray}
&&\text{Prob}[{\bf X}] \propto \exp\left[-\sum_{i=1}^{N_t} \sum_{k=1}^N
\frac{1}{4 D_k \Delta t} \right. \nonumber\\
&&\cdot \left. \left({\bf x}^k(i+1)-{\bf x}^k(i) + \frac{\Delta t ~D_k}{k_B T} \nabla U[{\bf X}(i)] \right)^2\right],
\label{prob}
\end{eqnarray}
where $N_t$ is the number of time steps in the trajectory. On the other hand, the paths which do not reach native state before time $t$ do not contribute to the
transition probability in Eq.~{\bf \ref{Pf}}.
The most probable ---or so-called \emph{dominant}--- reaction pathways are those which minimize the exponent in Eq. {\bf \ref{prob}}. In principle, these may be found by numerically
relaxing the effective action functional~\cite{DRP1, Adib}
\begin{eqnarray}
S_{eff}[{\bf X}] = \Delta t~\sum_{i=1}^{N_t}~\left[\sum_{k=1}^N \frac{({\bf x}_k(i+1)-{\bf x}_k(i))^2}{4 D_k \Delta t^2} + V_{eff}[{\bf X}(i)]\right],
\label{Seff}
\end{eqnarray}
where $V_{eff} ({\bf X})$ is the so-called effective potential, and reads
\begin{equation}
V_{eff} ({\bf X}) = \frac{1}{4 (k_B T)^2} \sum_k D_k~\left(~ \left| {\bf \nabla}_k~U({\bf X}) \right|^2 - 2~k_B
T\; \nabla_k^2~U({\bf X}) \ \right).
\label{Veff}
\end{equation}
In practice, for a protein folding transition, directly minimizing the effective action in Eq.~{\bf \ref{Seff}} is unfeasible, since at least $10^4-10^5$ time steps are needed to describe a single
folding event. On the other hand, for any fixed pair of native and denatured configurations the dominant paths can be equivalently found by
minimizing an effective Hamilton-Jacobi (HJ) action in the form~\cite{DRP1,DRP0}
\begin{equation}
S_{HJ} =\sum_{i=1}~\Delta l_{i,i+1} \sqrt{\frac{1}{D}\left(E_{eff} + V_{eff}[{\bf X}(i)]\right)},
\label{SHJ}
\end{equation}
where and $\Delta l_{i+1, i}=\sqrt{({\bf X}(i+1)-{\bf X}(i))^2}$ represents the elementary
displacement in configuration space, and for sake of clarity, we have assumed that all atoms have the same diffusion coefficient.
The parameter $E_{eff}$ determines the time at which any given frame $l$ of the path is visited:
\begin{eqnarray}
t(l) = \sum_{i\le l} \Delta l_{i, i+1} ~\hspace{0.1cm}[4 D(E_{eff}+V_{eff}[{\bf X}(i)]~)]^{-1/2}.
\label{time}
\end{eqnarray}
Hence, by adopting the HJ formulation of Eq.~{\bf \ref{SHJ}}, it is possible to replace
the time discretization with the discretization of the curvilinear abscissa $l$, which measures the
Euclidean distance covered in configuration space during the reaction~\cite{DRP0}. This way, the problem of the decoupling of time scales is bypassed.
As a result, only about $10^2$ frames are usually sufficient to provide a convergent representation of a trajectory.
On the other hand, the HJ formalism requires to perform an optimization in the space of reactive pathways of a functional, which can take complex values, which is in general
a complicated task.
The DRP approach displays important similarities with the SDEL (Stochastic Difference Equation in Length) method developed by Elber and co-workers~\cite{method4}.
In particular, while the DRP is based on minimizing the effective HJ action $S_{DRP} = 1/\sqrt{D}\int dl \sqrt{V_{eff}[{\bf X}(l)]+E_{eff}}$, in SDEL the folding trajectories
are obtained by {\it extremizing} the {\it physical} HJ action $S_{SDEL} = \int dl \sqrt{U[{\bf X}(l)]-E}$, where $U(x)$ is the potential energy and $E$ is the conserved
total mechanical energy.
The SDEL algorithm is presently implemented in the MOIL software for molecular modeling~\cite{MOIL}. In this code, the protein folding trajectories are obtained by means of
simulating annealing relaxation, starting from an initial trial trajectory --see e.g. Ref. \cite{Elberproteins}---.
The DRP algorithm is now implemented in DOLOMIT (Dynamics Of muLti-scale mOdels of Molecular transITions), a program under development at the Interdisciplinary Laboratory for
Computational Science (Trento). This code uses the MD-based algorithm described in the next section to identify the dominant path.
\subsection{Exploration of the Path Space}
The reliability of the DRP approach in investigating the protein folding transition crucially depends on the efficiency of the algorithm used to find optimum paths.
In the analysis of conformational \cite{DRP2, QDRP2} or chemical~\cite{QDRP1} reactions of relatively small molecules, dominant paths can be found by directly
optimizing the HJ action in Eq.~{\bf \ref{SHJ}}, e.g. using
simulated annealing or gradient-based methods. The DRP calculations for protein folding obtained this way
have been extensively tested using reduced models in which the relevant degrees of freedom are
individual aminoacids and the energy landscape was relatively smooth~\cite{testDRP1, testDRP2}. Unfortunately,
when moving from a coarse-grained to an atomistic description, a large amount of frustration is introduced and the energy surface of the model
becomes much rougher. Consequently, the relaxation algorithms adopted in our previous calculations
were found to provide a poor exploration of the space of folding paths in an all-atom calculation.
In order to overcome this problem, we have used a biased MD algorithm to efficiently produce a large ensemble of paths, starting from a given
denatured configuration and reaching the native state \cite{ratchetMD2,ratchetMD3, method3}.
In particular, the so-called \emph{ratchet-and-pawl MD} (rMD) algorithm \cite{method3} exploits the spontaneous fluctuations of the system along a specific collective coordinate (CC),
towards its native configuration.
This is done by introducing a time-dependent bias potential $V_R({\bf X},t)$, whose purpose is to make it very unlikely for the system to evolve back to previously
visited values of the CC. On the other hand, this bias exerts no work on the system when it spontaneously proceeds towards the native state.
We emphasize that this approach is quite different from the one used in \emph{steered}-MD~\cite{steeredMD}, where an external force is continuously
applied to the system, in order to drive it towards the desired state.
Following the work of Ref. \cite{method3} we chose a CC $z(t)$, which defines the distance between the contact map
in the instantaneous configuration ${\bf X}(t)$ from the contact
map in the native configuration ${\bf X}^{\text{native}}$. This way, the energy optimized native configuration has by definition $z=0$. On the other hand, a bias on $z$ does not force nor lock any
specific contact, but only imposes
a (quasi)~monotonic behavior of the \emph{total number} of native contacts.
In particular, the biasing potential introduced in Ref. \cite{method3} is defined as
\begin{eqnarray}
V_R({\bf X}, t) = \left\{
\begin{array}{lr r}
\frac{k_R}{2} ( z[{\bf X} (t)] - z_m(t) )^2, &\text{for}&\quad z[{\bf X}(t)] > z_m(t)\\
0, &\text{for}&\quad z[{\bf X}(t)] \leq z_m(t).\\
\end{array}
\right.\vspace{2mm}
\label{VR}
\end{eqnarray}
In these equations, $z_m(t)$ is the minimum value assumed by the collective variable $z$ along the rMD trajectory, up to time~$t$
The value of the collective variable $z$ in the instantaneous configuration $X(t)$ is defined as:
\begin{equation}
z[X(t)] \equiv \sum_{i,j}^{N} [ C_{ij}[{\bf X}(t)] - C_{ij}({\bf X}^{\text{native}}) ]^2.
\end{equation}
The entries of the contact map C$_{ij}$ are chosen to interpolate smoothly between 0 and 1, depending on the relative distance of the
residues $i$ and $j$:
\begin{equation}
C_{ij}({\bf X}) = \frac { 1-(\frac{ r_{ij} }{ r_0 })^6 } { 1-(\frac{ r_{ij} }{ r_0 })^{10} },
\end{equation}
where r${_0}$=7.5~\AA~ is a fixed reference distance.
The variable $z_m$(t) is updated
only when the system visits a configuration with a smaller value of the CC, i.e. any time $z[{\bf X}(t+\delta t)]< z_m(t)$.
The value of the spring constant $k_R$ in the ratchet potential ---see Eq. {\bf \ref{VR}}--- controls the amount of bias introduced by the ratchet algorithm.
At very low values of $k_R$, the rMD trajectories are minimally biased. However, in this limit the system performs the
first folding transition on typical time intervals comparable with the folding time $t\sim 1/k_f$ ---where
$k_f$ is the folding rate---.
On the other hand, the DRP method is based on the transition probability given Eq. {\bf \ref{Pf}}, where the time interval $t$ is of
the order of the transition path time $t_{TPT}\ll 1/k_f$.
Hence, DRP simulations based on ratchet simulations in which $k_R$ is chosen too small are computationally very expensive, since most paths do not reach the
native state within the time interval $t$ entering Eq. {\bf \ref{Pf}}, hence do not contribute to the transition probability.
In the opposite high $k_R$ limit, the bias force becomes comparable with the physical internal forces acting on the atoms and the dynamics is affected by a significant bias. In this regime, if the ratcheting coordinate is not optimal, the system is driven into relatively large free energy regions. The unbiased statistical weight given by Eq. {\bf \ref{prob}}
penalizes these trajectories.
In the extremely high $k_R$ limit, the bias towards forming native contacts is very large, and breaking a native contact which is present in a (partially) misfolded configuration becomes extremely unlikely. As a result, also rMD simulations in which $k_R$ is chosen too large have a low folding yield, since the trajectories have a large probability to be trapped in misfolded states.
This algorithm allows to efficiently generate a large number of trajectories starting from the same configuration and reaching the native state, hence it can be used to explore the
folding path space. Eq. {\bf \ref{prob}} provides a rigorous way to score such trial trajectories, i.e.
to evaluate the probability for each of them to be realized in an \emph{unbiased} overdamped Langevin dynamics simulation.
In particular, the best estimate for the dominant folding pathway is the one with the smallest Onsager-Machlup action.
The path may then be used as a starting point for a further refinement based on a local relaxation of the HJ action given by
Eq.~{\bf \ref{SHJ}}, performed by means of the optimization algorithms described in our previous work (see e.g. Ref. \cite{QDRP2, QDRP3}).
The second refinement step is computationally very expensive, requiring several thousands of CPU hours for each dominant trajectory. However,
by performing a number of test simulations, we have found that it produces only very small rearrangements of the chain, mostly filtering out small thermal fluctuations
(see e.g. Fig.\ref{Fig4}).
Some comments on the minimization procedure described above are in order: first, we emphasize that since the dynamics generated by the rMD algorithm has infinite memory, it
violates the detailed-balance condition and the time intervals computed from the biased trajectories have no direct physical meaning.
On the other hand, once a minimum of the HJ action has been identified, it is in principle possible to reconstruct the physical time at which each frame is visited, by means of
Eq.~{\bf \ref{time}}.
As long as one is concerned mostly with the global qualitative aspects of the folding mechanism, the expensive refinement session of the DRP calculation may be dropped.
This allows us to reduce the total computational time required to perform the analysis by several orders of magnitude.
On the other hand, we are not in a condition to provide a reliable estimate of the
time interval that the protein takes to complete a folding reaction.
We stress that the present DRP approach is expected to give unreliable results if the sampling of the trial trajectories is too limited or if the CC
adopted in the bias potential of the rMD algorithm was a bad reaction coordinate.
A number of recent studies have shown that the total number of native contacts represents reasonable reaction coordinate for the folding of small globular proteins
~\cite{ reactioncoord2, reactioncoord1,reactioncoord3}.
\begin{table*}[th]
\caption{Summary of the simulation details}
\begin{tabular*}{\hsize}
{@{\extracolsep{\fill}}rrrrr}
\hline
\multicolumn1c{Unfold. MD steps} &
\multicolumn1c{Therm. MD steps } &
\multicolumn1c{Fold. rMD steps} &
\multicolumn1c{Denatured} &
\multicolumn1c{Rejected}
\cr
\multicolumn1c{(T= 1600 K)} &
\multicolumn1c{(T= 300 K)} &
\multicolumn1c{ (T= 300 K)} &
\multicolumn1c{initial conditions} &
\cr
\multicolumn1c{$50 \times 10^3$\tablenote{In all MD and rMD simulations the integration time step has been chosen to be 1 fs.}} &
\multicolumn1c{$100 \times 10^3$} &
\multicolumn1c{$50 \times 10^3$} &
\multicolumn1c{44} & \multicolumn1c{18}
\cr
\hline
\end{tabular*}
\label{table}
\end{table*}
\subsection{Computational Procedures and Simulation Details}
The atomistic DRP calculations were performed using DOLOMIT. The code calls a
librarized version of GROMACS 4.5.2~\cite{GRO4} to calculate the molecular potential energy and its gradient.
We defined the native state as the set of conformations with a RMSD to the crystal structure of the $C_\alpha$ in the hairpins smaller than 3.5 \AA.
A configuration was considered
denatured if the RMSD to native of both hairpins was larger than 6\AA.
The stability of the native state within the present force field was checked by running 12 unbiased 2 ns-long MD simulations at the room temperature (300 K). In all such trajectories
the protein remained in the native state. We then generated 44 independent initial conditions, by running a 50~ps MD at 1600~K, starting from the
energy minimized native state, followed by a 100 ps relaxation at 300~K. The time step employed in all the simulations was 1 fs.
From each of the 24 starting configurations we performed 96 independent runs, each consisting of 50000 rMD steps. For each of the remaining 20 initial conditions, the number of trial paths was limited to 48.
In our simulations, we have chosen the ratchet spring constant
$k_R=0.02$~kcal/mol, which represents a reasonable compromise between keeping a high computational efficiency for the path exploration
and introducing a small bias. Indeed, using this value, the modulus of the biasing force is always at least one order of magnitude smaller
than the norm of internal forces.
We observed that 5 of the 44 initial conditions did not correspond to denatured states, hence they were rejected.
In addition, in 13 of the remaining 39 sets, more than $80\%$ of the trial trajectories did not reach the native state within the simulation time. In these cases
the exploration of the path space was limited to very few trial trajectories, so the corresponding dominant paths were discarded.
For the remaining 26 sets of paths, we identified the most probable by computing the OM action of Eq. {\bf \ref{prob}}.
The complete set of atomistic simulations required less than two days of calculation on 48 CPU's.
It is important to study to what extent DRP results obtained this way depend on the value of the bias constant $k_R$ adopted in the rMD
simulations.
In Fig. \ref{Fig5} we plot the dominant reaction pathway obtained starting from the same initial condition, using different values of $k_R$ which span over almost two orders of
magnitude. We see that in most simulations the folding occurs through the same qualitative pathway, in which the first hairpin forms before the second begins to fold. Only
in one case ---for a low value of the coupling constant $k_R$--- we find that the protein travels across the denatured state before taking a different pathway to the native state,
in which the order of formation of the hairpins is reversed.
Such a trajectory spends a much longer fraction of rMD steps in exploring the denatured state and initiates the transition from a very different configuration.
It is interesting to note that the most probable path turns out not to be the one with the lowest ratchet constant.
In particular, the trajectory taking the second pathway (labelled with $K_1$ in Fig. \ref{Fig5}) is among those with the lowest statistical weight. This is because the OM action
tends to penalize paths which travel long Euclidean distances in configuration space. This result
illustrates that choosing a low ratchet constant $k_R$ produces an enhanced exploration of the denatured state, but does not necessarily lead to a more
efficient identification of the most probable trajectories connecting given boundary condition at fixed times.
Once a dominant path has been found, it is relatively straightforward to identify
the configuration which belongs to the transition state ensemble. This can be done by finding the frame ${\bf X}_{TS}$ in the trajectory such that
the probability to reach the native state is equal to that of going back to the denatured configuration~\cite{DRP2}:
\begin{eqnarray}
\label{TS}
\frac{\text{Prob}[{\bf X}_{TS}\to \text{Unfolded}]}{\text{Prob}[{\bf X}_{TS}\to \text{Native}]} \simeq \frac{e^{-S_{OM}({\bf X}_{TS}\to {\bf X}_D)}}{e^{-S_{OM}({\bf X}_{TS}\to {\bf X}_N)}}=1
\end{eqnarray}
In this equation ${\bf X}_N$ and ${\bf X}_D$ are the first native and denatured configurations visited along the dominant path, starting from ${\bf X}_{TS}$.
In order to locate them, we need to only take into account the ``reactive'' part of the path, that is the one which leaves the denatured state and, without recrossing,
goes straight to the native. To satisfy this requirement, we considered the total RMSD versus frame index curve. The typical trend of this curve for most of the dominant trajectories is
shown in the lower panel of Fig.~\ref{Fig6}: it consists
in an initial plateau, followed by a rather steep fall, and then by another flat region, where
the system oscillates in the native state. The reactive part of the path was identified with the region of steep fall in this curve. In particular,
the beginning of the transition was set to the frame at which the derivative of the total RMSD curve changes sign, from positive to negative.
The simulations of the equilibrium properties of the WW-domain in the coarse-grained models were
performed using a Monte Carlo algorithm based on a combination of Cartesian, crankshaft~\cite{crankshaftMC} and pivot~\cite{pivot} moves.
The free energy as a function of an arbitrary set of reaction coordinates (potential of mean force) was obtained from the frequency histogram
calculated from long MC trajectories.
\section{Results and Discussion}
In the upper panel of Fig.\ref{Fig6} we show our set of atomistic dominant folding trajectories, projected onto the plane defined by the
RMSD to the native structure
of the C$_{\alpha}$ atoms in residues $8-23$ (hairpin~1) and $17-30$ (hairpin~2).
Two distinct folding pathways which differ by the order of formation of the hairpins can be clearly identified: in about half of the computed dominant
folding pathways hairpin~1 consistently folds before hairpin 2.
In this channel, we find the transition state is located at the ``turn" of the paths, i.e. is formed by configurations in which the hairpin 1 is folded
while hairpin 2 is largely unstructured (see pathway 1 in Fig. \ref{Fig2}). This is the mechanism predominantly found in the simulation of Ref.~\cite{theWW5}, performed using the same force field, albeit in explicit solvent.
In about half of the computed dominant paths, we observe that the two hairpins form in the reversed order. In this channel, the transition state is formed by the
configurations in which hairpin 2 is
folded, while hairpin 1 is unstructured (see pathway 2 in Fig.~\ref{Fig2}).
An important point to emphasize is that this clean two-channel folding mechanism is a prediction of the DRP approach which could not be obtained if we analyzed directly
the full set of (biased) trajectories obtained from the rMD algorithm, without scoring their probability.
For example, Fig.~\ref{Fig7} shows that not all the rMD trajectories computed starting from a given initial condition follow one of the two folding pathways discussed above.
Indeed, many of them
involve a simultaneous formation of native contacts in both hairpins. A clear prediction of the DRP formalism is that folding events in which the hairpins form simultaneously are much less
frequent than those in which the two secondary structures forms in sequence.
Another result emerging from our DRP calculation is the existence of a correlation between the structure of the initial conditions from which the transition is initiated
and the pathway taken to fold: if at the beginning of the transition
the first hairpin has a RMSD smaller than the second hairpin, then the first pathway is most likely chosen.
In the opposite case, i.e. when the second hairpin has a smaller RMSD to native than the first, then the second pathway is generally preferred.
In order to further support these results and gain insight into the folding mechanism, we have performed simulations in an entirely different approach, i.e. by computing equilibrium properties
using the coarse-grained models described in the Methods section. In Fig. \ref{Fig3} we show the free energy landscape at the 300~K, as a function of the RMSD
to native of the two hairpins for the two models, which differ by the presence of non-native interactions.
In both cases, we observe the existence of two valleys in the free energy landscape, which correspond to the two folding pathways discussed above. Remarkably, the same structure for this
free energy map was obtained by Ferrara and co-workers for the 20-residue peptide beta3s ---which shares the same native topology of WW domains \cite{ferrara}---
by means of equilibrium atomistic simulations based on the CHARMM force field, in implicit solvent.
We emphasize that this set of qualitatively consistent results has been obtained using different algorithms (in equilibrium and nonequilibrium conditions) and different theoretical models
(atomistic and coarse-grained). The fact that models with and without non-native interactions give very similar free energy landscapes suggests
that the structure of the two folding pathways of WW domains is mostly shaped by native interactions.
\subsection{Comparison with experimental data on folding kinetics}
In general, all experimental data on folding kinetics of WW domains indicate that the formation of the first hairpin is the main rate limiting step~\cite{expWW1, expWW2, expWW3}.
In particular, the $\phi$-values measured by J\"ager and co-workers display a clear peak in the region associated with hairpin 1, but significant
$\phi$--values were reported also for residues in the sequence region relative to hairpin 2~\cite{expWW3}. This fact indicates that the folding of the latter structure has some rate limiting effect.
In addition, it was found that the $\phi$--values in the region of the second hairpin grow with temperature, while those in the region of the first hairpin decrease.
This implies that, at higher and higher temperatures, the second hairpin plays an increasing role in the folding mechanism.
An analysis based on $\phi-$values alone does not permit to fully characterize the folding mechanism. In particular, it cannot distinguish between a single-channel folding
mechanism in which native contacts in the
two hairpins form simultaneously and a multiple-channel folding mechanism in which the reaction rate in each channel is limited by the folding of one of the hairpins.
In Ref. \cite{weikl} Weikl has shown that the full body of existing $\phi$--value data (taken from Ref.s \cite{expWW2, expWW3}) can be consistently and quantitatively explained by a simple kinetic model in which the folding of WW domains occurs through alternative channels,
which correspond to the two pathways found in our DRP simulations. From a global fit of the experimental data, the author concluded that the relative probability of the
first folding pathway for FBP and Pin 1 WW domain are $77\% \pm5\%$ and $67 \pm 5\%$, respectively.
Let us now discuss the relative statistical weight of the two folding pathways. To this goal we need to estimate and compare
the reaction rates in the two channels. The formalism for evaluating reaction rates in the DRP approach was developed in detail in Ref. {\bf \cite{DRP4}},
where it was shown that this method reproduces Kramers theory in the low-temperature regime. Applying that formalism, the ratio of the folding rates in the two channels reads
\begin{eqnarray}
\label{rate}
\frac{k_1}{k_2} \simeq \frac{\kappa^1_0}{\kappa^2_0}~e^{-\beta~(G_{TS_1}-G_{TS_2})},
\end{eqnarray}
where the label 1 (2) identifies the channel in which hairpin 1 (2) folds first.
In Eq.~{\bf \ref{rate}}, the exponent contains the difference of the free-energies of the two transition states, defined from the dominant trajectories according to the commitment analysis
described in the Methods section and in Ref.~ \cite{DRP2}.
In particular, one has
\begin{eqnarray}
\label{Gi}
e^{-\beta G(TS_i)} \equiv \int d{\bf X}~ e^{-\beta U({\bf X})} \delta\left[({\bf X} -{\bf X}_{TS}^i) \cdot \hat n_{TS^i}\right]\qquad (i=1,2),
\end{eqnarray}
where ${\bf X}^i_{TS}$ is a point of the transition state which is visited by a typical dominant path in the $i-$th reaction channel and $\hat n_{TS^i}$ is a
versor tangent to the dominant path at ${\bf X}_{TS}^i$. Using Eq.~{\bf \ref{TS}} to identify the transition states, we have found that the two partition functions defined in Eq.~{\bf\ref{Gi}} are dominated by configurations in which one of the hairpin is
fully formed while the other is still completely unstructured (see Fig. \ref{Fig6}). The average location of the computed transition states in the plane defined by the
RMSD to native of the two hairpins is highlighted in the lower panel of Fig. \ref{Fig3} and can be identified with the endpoints of the two low free energy valleys.
The coefficients $\kappa_0^1$ and $\kappa_0^2$ in the prefactor of Eq.~{\bf \ref{rate}} are defined in terms of quantities which can be calculated from the dominant paths --- see Ref.\cite{DRP4} for details---. These terms
estimate the average flux of reactive trajectories
across the iso-committor dividing surface, including the contributions from small thermal fluctuations around the dominant paths. Unfortunately, evaluating $\kappa_0^1$ and $\kappa_0^2$
necessarily requires to perform the computationally expensive local optimization of the HJ action.
However, if the reaction is thermally activated, the ratio of rates $k_1/k_2$ is mostly controlled by the exponential contribution. Hence, we can consider the Arrhenius
approximation
\begin{eqnarray}\label{ratearrhenius}
\frac{k_1}{k_2} \simeq e^{-\beta~(G_{TS_1}-G_{TS_2})}.
\end{eqnarray}
We emphasize that in this approach the DRP information about the nonequilibrium reactive dynamics is used to define the two transition states.
On the other hand, the numerical value of the free energy difference may be obtained from equilibrium techniques, e.g. by sampling the integrals in Eq.~{\bf \ref{Gi}} by means of computationally very expensive umbrella sampling or
meta-dynamics~{\bf \cite{Metadinamica}} atomistic calculations.
In this first exploratory application of the DRP formalism to a realistic protein folding reaction we choose to perform a much rougher estimate which relies on two main approximations. First, we identify the difference $(G_{TS_1}-G_{TS_2})$ in Eq.~{\bf \ref{ratearrhenius}} with the difference of the free energy in the two shaded regions
of the energy landscape shown in Fig. \ref{Fig3}. The centers of these regions represent the average location of the configurations in the two transition states TS1 and TS2 obtained from DRP simulations, projected onto the plane selected by the RMSD to native of the two hairpins. The sizes of the shaded area represents the errors on the average location of the transition states on this plane, estimated from the standard deviation.
The second assumption of our model is
that such a free energy difference is driven by the balance between energy gain and entropy loss associated to the formation of \emph{native} contacts in the two hairpins. This native-centric standpoint is supported in part by the fact that
free energy landscapes computed in different models with and without non-native interactions are found to be very similar, as it is clear from comparing the panels in Fig.~\ref{Fig3}.
Hence, to estimate
$G_{TS_1}-G_{TS_2}$ we used the G$\overline{o}$-type model described in the Methods section.
It is important to emphasize that we are not computing the rate directly from a transition state theory formulated in the coarse-grained model,
but we are using it only to estimate a free energy difference.
This way, we obtained an estimate $k_1/k_2 \simeq 2.3$ which corresponds to a relative weight of the first folding channel of $70\%$ and $30\%$.
We stress that, such a simple calculation should be considered only a rough estimate. It indicates that the two channels have more or less comparable weight and that the first channel is the most
probable, in qualitative agreement with experimental results and with the simulations of Shaw and co-workers.
This simple scheme enables us to address the question of the dependence of the relative weights of the two channels on the temperature.
Repeating the calculation at a higher temperature of $380$~K --- assuming that the structure of the transition states is not significantly modified--- we find $k_1/k_2 \simeq 1.6$, which corresponds to a branching ratio
of channel 1 of about $60\%$. Hence, the rate limiting role of the second channel grows with temperature, in qualitative agreement with experimental kinetic data.
This fact can be understood as follows. The folding of one of the hairpins generates an entropy loss proportional to the number $n$ of native contacts formed.
More precisely, assuming that the protein in the unfolded state is completely denatured then the entropy loss in the TS of the first (second) channel is $\Delta S_{TS1 (TS2)}= n_1 (n_2) s$, where $s$ is the entropy loss for locking one residue in its native conformation, while $n_1=15$ and $n_2=13$ are the numbers of residues in the two hairpins of Fip35.
The transition state in the first channel involves forming a longer hairpin, hence reaching it involves a larger entropy loss (but also larger gain of native energy).
The role of the entropy loss relative to the energy gain in forming the hairpins grows with temperature, hence
disfavoring the first folding channel relative to the second.
\section{Conclusions}
In this work we have studied the folding mechanism of the WW domain Fip35
in equilibrium and nonequilibrium conditions, using an atomistically detailed force field and coarse-grained models based on knowledge-based potentials.
The obtained folding trajectories are not heterogeneous, but rather suggest that the folding proceeds through two dominant channels, defined by a hierarchical order of hairpin formation.
In the most probable channel at room temperature, the first hairpin is almost fully formed before the second hairpin starts to fold.
In the alternative route, the order of hairpin formation is reversed.
Our simulations also suggest that the choice of the folding pathway is correlated with the structure of the denatured configuration from which the peptide initiates the folding reaction.
Our results are compatible with both Weikl's analysis of kinetic data and Krivov's analysis of equilibrium MD simulations.
The most important result of this work is to show that, using the DRP approach, it is possible to characterize at least the main qualitative aspects of the folding mechanism
at an extremely modest computational cost, in the range of few hundreds of CPU hours. Such a level of computational efficiency opens the door to the investigation of the folding pathways of a large number of single-domain proteins, with sizes significantly larger than that of the small domain studied in the present work.
\begin{acknowledgments}
The DRP approach was developed in collaboration with H. Orland, M. Sega and F. Pederiva.
We thank G. Tiana and C. Camilloni for sharing important details of their implementation of the ratched-MD algorithm and S. Piana-Agostinetti for providing us with details on the results
of the MD simulations of Ref. \cite{theWW5}.
All the authors are members of the Interdisciplinary Laboratory for Computational Science (LISC), a joint venture of Trento University and the Bruno Kessler Foundation.
SaB and T\v{S} are supported by the Provincia Autonoma di Trento, through the AuroraScience project.
Simulations were performed on the AURORA supercomputer located at the LISC (Trento) and partially on the TITANE cluster, which was kindly made available by the IPhT of CEA-Saclay.
\end{acknowledgments}
|
\p@rtnewp@ge\addvspace\p@rtbeforeskip\@afterindentfalse\secdef\@part\@spart{\p@rtnewp@ge\addvspace\p@rtbeforeskip\@afterindentfalse\secdef\@part\@spart}
\def\@part[#1]#2{\ifnum \c@secnumdepth >-1\relax
\refstepcounter{part}
\def\@tempa{\addcontentsline{toc}{part}} %
\expandafter\@tempa\expandafter{\thepart
\hspace{1em}#1}\else
\addcontentsline{toc}{part}{#1}\fi
{\p@rtstyl@
\ifnum \c@secnumdepth >-1\relax
{\partintr@styl@Part\ \thepart
\partintr@dot}\partintr@sep\nobreak
\fi
#2\partd@t\markboth{}{}\par}
\nobreak
\vskip\p@rtafterskip
\@afterheading
}
\def\@spart#1{{\p@rtcentering\p@rtstyl@
#1\partd@t\par}
\nobreak
\vskip\p@rtafterskip
\@afterheading
}
\newif\ifsection@ftind
\newif\ifsection@ftpar
\def\sectionbeforeskip#1{\def\s@ctbeforeskip{#1}}
\def\sectionstyle#1{\def\s@ctstyl@{#1}}
\def\sectiondot#1{\def\sectiond@t{#1}}
\def\sectionafterskip#1{\def\s@ctafterskip{#1}}
\def\sectionintrostyle#1{\def\sectionintr@styl@{#1}}
\def\sectionintro#1{\def\sectionintr@{#1}}
\def\sectionintrodot#1{\def\sectionintr@dot{#1}}
\def\sectionintrosep#1{\def\sectionintr@sep{#1}}
\def\def\s@ctind{\parindent}{\def\s@ctind{\parindent}}
\def\def\s@ctind{\z@}{\def\s@ctind{\z@}}
\def\@startsection{section}{1}{\s@ctind@ftindtrue{\@startsection{section}{1}{\s@ctind@ftindtrue}
\def\@startsection{section}{1}{\s@ctind@ftindfalse{\@startsection{section}{1}{\s@ctind@ftindfalse}
\def\@startsection{section}{1}{\s@ctind@ftpartrue{\@startsection{section}{1}{\s@ctind@ftpartrue}
\def\@startsection{section}{1}{\s@ctind@ftparfalse{\@startsection{section}{1}{\s@ctind@ftparfalse}
\newif\ifsubsection@ftind
\newif\ifsubsection@ftpar
\def\subsectionbeforeskip#1{\def\ss@ctbeforeskip{#1}}
\def\subsectionstyle#1{\def\ss@ctstyl@{#1}}
\def\subsectiondot#1{\def\subsectiond@t{#1}}
\def\subsectionafterskip#1{\def\ss@ctafterskip{#1}}
\def\subsectionintrostyle#1{\def\subsectionintr@styl@{#1}}
\def\subsectionintro#1{\def\subsectionintr@{#1}}
\def\subsectionintrodot#1{\def\subsectionintr@dot{#1}}
\def\subsectionintrosep#1{\def\subsectionintr@sep{#1}}
\def\def\ss@ctind{\parindent}{\def\ss@ctind{\parindent}}
\def\def\ss@ctind{\z@}{\def\ss@ctind{\z@}}
\def\@startsection{subsection}{2}{\ss@ctind@ftindtrue{\@startsection{subsection}{2}{\ss@ctind@ftindtrue}
\def\@startsection{subsection}{2}{\ss@ctind@ftindfalse{\@startsection{subsection}{2}{\ss@ctind@ftindfalse}
\def\@startsection{subsection}{2}{\ss@ctind@ftpartrue{\@startsection{subsection}{2}{\ss@ctind@ftpartrue}
\def\@startsection{subsection}{2}{\ss@ctind@ftparfalse{\@startsection{subsection}{2}{\ss@ctind@ftparfalse}
\newif\ifsubsubsection@ftind
\newif\ifsubsubsection@ftpar
\def\subsubsectionbeforeskip#1{\def\sss@ctbeforeskip{#1}}
\def\subsubsectionstyle#1{\def\sss@ctstyl@{#1}}
\def\subsubsectiondot#1{\def\subsubsectiond@t{#1}}
\def\subsubsectionafterskip#1{\def\sss@ctafterskip{#1}}
\def\subsubsectionintrostyle#1{\def\subsubsectionintr@styl@{#1}}
\def\subsubsectionintro#1{\def\subsubsectionintr@{#1}}
\def\subsubsectionintrodot#1{\def\subsubsectionintr@dot{#1}}
\def\subsubsectionintrosep#1{\def\subsubsectionintr@sep{#1}}
\def\def\sss@ctind{\parindent}{\def\sss@ctind{\parindent}}
\def\def\sss@ctind{\z@}{\def\sss@ctind{\z@}}
\def\@startsection{subsubsection}{3}{\sss@ctind@ftindtrue{\@startsection{subsubsection}{3}{\sss@ctind@ftindtrue}
\def\@startsection{subsubsection}{3}{\sss@ctind@ftindfalse{\@startsection{subsubsection}{3}{\sss@ctind@ftindfalse}
\def\@startsection{subsubsection}{3}{\sss@ctind@ftpartrue{\@startsection{subsubsection}{3}{\sss@ctind@ftpartrue}
\def\@startsection{subsubsection}{3}{\sss@ctind@ftparfalse{\@startsection{subsubsection}{3}{\sss@ctind@ftparfalse}
\newif\ifparagraph@ftind
\newif\ifparagraph@ftpar
\def\paragraphbeforeskip#1{\def\p@rbeforeskip{#1}}
\def\paragraphstyle#1{\def\p@rstyl@{#1}}
\def\paragraphdot#1{\def\paragraphd@t{#1}}
\def\paragraphafterskip#1{\def\p@rafterskip{#1}}
\def\paragraphintrostyle#1{\def\paragraphintr@styl@{#1}}
\def\paragraphintro#1{\def\paragraphintr@{#1}}
\def\paragraphintrodot#1{\def\paragraphintr@dot{#1}}
\def\paragraphintrosep#1{\def\paragraphintr@sep{#1}}
\def\def\p@rind{\parindent}{\def\p@rind{\parindent}}
\def\def\p@rind{\z@}{\def\p@rind{\z@}}
\def\@startsection{paragraph}{4}{\p@rind@ftindtrue{\@startsection{paragraph}{4}{\p@rind@ftindtrue}
\def\@startsection{paragraph}{4}{\p@rind@ftindfalse{\@startsection{paragraph}{4}{\p@rind@ftindfalse}
\def\@startsection{paragraph}{4}{\p@rind@ftpartrue{\@startsection{paragraph}{4}{\p@rind@ftpartrue}
\def\@startsection{paragraph}{4}{\p@rind@ftparfalse{\@startsection{paragraph}{4}{\p@rind@ftparfalse}
\newif\ifsubparagraph@ftind
\newif\ifsubparagraph@ftpar
\def\subparagraphbeforeskip#1{\def\sp@rbeforeskip{#1}}
\def\subparagraphstyle#1{\def\sp@rstyl@{#1}}
\def\subparagraphdot#1{\def\subparagraphd@t{#1}}
\def\subparagraphafterskip#1{\def\sp@rafterskip{#1}}
\def\subparagraphintrostyle#1{\def\subparagraphintr@styl@{#1}}
\def\subparagraphintro#1{\def\subparagraphintr@{#1}}
\def\subparagraphintrodot#1{\def\subparagraphintr@dot{#1}}
\def\subparagraphintrosep#1{\def\subparagraphintr@sep{#1}}
\def\def\sp@rind{\parindent}{\def\sp@rind{\parindent}}
\def\def\sp@rind{\z@}{\def\sp@rind{\z@}}
\def\@startsection{subparagraph}{4}{\sp@rind@ftindtrue{\@startsection{subparagraph}{4}{\sp@rind@ftindtrue}
\def\@startsection{subparagraph}{4}{\sp@rind@ftindfalse{\@startsection{subparagraph}{4}{\sp@rind@ftindfalse}
\def\@startsection{subparagraph}{4}{\sp@rind@ftpartrue{\@startsection{subparagraph}{4}{\sp@rind@ftpartrue}
\def\@startsection{subparagraph}{4}{\sp@rind@ftparfalse{\@startsection{subparagraph}{4}{\sp@rind@ftparfalse}
\sectionbeforeskip{\bigskipamount}
\sectionstyle{\large\texbf\boldmath}
\sectiondot{}
\sectionafterskip{.5\bigskipamount}
\sectionintrostyle{}
\sectionintro{}
\sectionintrodot{.}
\sectionintrosep{1.25ex}
\def\s@ctind{\z@}
\@startsection{section}{1}{\s@ctind@ftindtrue
\@startsection{section}{1}{\s@ctind@ftpartrue
\subsectionbeforeskip{.8\bigskipamount}
\subsectionstyle{\normalsize\texbf\boldmath}
\subsectiondot{}
\subsectionafterskip{.4\bigskipamount}
\subsectionintrostyle{}
\subsectionintro{}
\subsectionintrodot{.}
\subsectionintrosep{1.25ex}
\def\ss@ctind{\z@}
\@startsection{subsection}{2}{\ss@ctind@ftindtrue
\@startsection{subsection}{2}{\ss@ctind@ftpartrue
\subsubsectionbeforeskip{.6\bigskipamount}
\subsubsectionstyle{\normalsize\texbf\boldmath}
\subsubsectiondot{}
\subsubsectionafterskip{.3\bigskipamount}
\subsubsectionintrostyle{}
\subsubsectionintro{}
\subsubsectionintrodot{.}
\subsubsectionintrosep{1.25ex}
\def\sss@ctind{\z@}
\@startsection{subsubsection}{3}{\sss@ctind@ftindtrue
\@startsection{subsubsection}{3}{\sss@ctind@ftpartrue
\paragraphbeforeskip{.5\bigskipamount}
\paragraphstyle{\normalsize\texbf\boldmath}
\paragraphdot{.}
\paragraphafterskip{1.25ex}
\paragraphintrostyle{}
\paragraphintro{}
\paragraphintrodot{.}
\paragraphintrosep{1.25ex}
\def\p@rind{\z@}
\@startsection{paragraph}{4}{\p@rind@ftindtrue
\@startsection{paragraph}{4}{\p@rind@ftparfalse
\subparagraphbeforeskip{.5\bigskipamount}
\subparagraphstyle{\normalsize\texbf\boldmath}
\subparagraphdot{.}
\subparagraphafterskip{1.25ex}
\subparagraphintrostyle{}
\subparagraphintro{}
\subparagraphintrodot{.}
\subparagraphintrosep{1.25ex}
\def\sp@rind{\parindent}
\@startsection{subparagraph}{4}{\sp@rind@ftindtrue
\@startsection{subparagraph}{4}{\sp@rind@ftparfalse
\let\@partoken\par
\long\def\@@gobble#1{}
\def\@ifnextchar\@partoken{\expandafter\ignorepar\@@gobble}{\ignorespaces}{\@ifnextchar\@partoken{\expandafter\@ifnextchar\@partoken{\expandafter\ignorepar\@@gobble}{\ignorespaces}\@@gobble}{\ignorespaces}}
\def\@startsection#1#2#3#4#5#6{
\@tempskipa #4\relax
\csname if#1@ftind\endcsname\@afterindenttrue\else\@afterindentfalse\fi
\advance\@tempskipa by\presection
\if@nobreak \everypar{}\else
\addpenalty{\@secpenalty}\addvspace{\@tempskipa}%
\allowbreak\vskip -\presection \fi \@ifstar
{\@ssect{#1}{#2}{#3}{#4}{#5}{#6}}{\@dblarg{\@sect{#1}{#2}{#3}{#4}{#5}{#6}}}}
\def\@sect#1#2#3#4#5#6[#7]#8{\def\object@type{#1}%
\ifnum #2>\c@secnumdepth\def\@svsec{}\def\@tempb{}%
\else\refstepcounter{#1}\def\@svsec{{\csname #1intr@styl@\endcsname%
{\csname #1intr@\endcsname}\csname the#1\endcsname%
\csname #1intr@dot\endcsname\kern\csname #1intr@sep\endcsname}}%
\edef\@tempb{\noexpand\numberline{\csname the#1\endcsname}}\fi%
\def\@tempa{\addcontentsline{toc}{#1}}%
\csname if#1@ftpar\endcsname%
\begingroup #6\relax%
\@hangfrom{\hskip #3\relax\@svsec}{\interlinepenalty \@M{#8}%
\csname #1d@t\endcsname\par}%
\endgroup%
\csname #1mark\endcsname{#7}%
\expandafter\@tempa\expandafter{\@tempb #7}%
\ifautolabel\label*{#8}\fi%
\else%
\def\@svsechd{#6\hskip #3\relax%
\@svsec{#8}%
\csname #1d@t\endcsname%
\csname #1mark\endcsname{#7}%
\expandafter\@tempa\expandafter{\@tempb #7}%
\ifautolabel\label*{#8}\fi}\fi%
\@xsect{#1}{#5}\@ifnextchar\@partoken{\expandafter\ignorepar\@@gobble}{\ignorespaces}}
\def\@ssect#1#2#3#4#5#6#7{%
\ifnum #2>\c@secnumdepth\def\@tempb{}\else \def\@tempb{\numberline{}}\fi%
\def\@tempa{\addcontentsline{toc}{s#1}}%
\csname if#1@ftpar\endcsname
\begingroup #6\relax
\@hangfrom{\hskip #3}{\interlinepenalty \@M{#7}%
\csname #1d@t\endcsname\par}%
\endgroup
\csname s#1mark\endcsname{#7}%
\ifstarredcontents\expandafter\@tempa\expandafter{\@tempb #7}\fi%
\ifautolabel\label*{#7}\fi%
\else%
\def\@svsechd{#6\hskip #3\relax{#7}%
\csname #1d@t\endcsname%
\csname s#1mark\endcsname{#7}%
\ifautolabel\label*{#7}\fi}\fi
\@xsect{#1}{#5}\@ifnextchar\@partoken{\expandafter\ignorepar\@@gobble}{\ignorespaces}}
\def\@xsect#1#2{
\csname if#1@ftpar\endcsname
\par \nobreak \vskip #2\relax \@afterheading
\else \global\@nobreakfalse \global\@noskipsectrue
\everypar{\if@noskipsec \global\@noskipsecfalse
\clubpenalty\@M \hskip -\parindent
\begingroup \@svsechd \endgroup \unskip
\hskip #2\relax
\else \clubpenalty \@clubpenalty
\everypar{}\fi}\fi\ignorespaces}
\def\@startsection{section}{1}{\s@ctind{\@startsection{section}{1}{\s@ctind}
{\s@ctbeforeskip}{\s@ctafterskip}{\s@ctstyl@}}
\def\@startsection{subsection}{2}{\ss@ctind{\@startsection{subsection}{2}{\ss@ctind}
{\ss@ctbeforeskip}{\ss@ctafterskip}{\ss@ctstyl@}}
\def\@startsection{subsubsection}{3}{\sss@ctind{\@startsection{subsubsection}{3}{\sss@ctind}
{\sss@ctbeforeskip}{\sss@ctafterskip}{\sss@ctstyl@}}
\def\@startsection{paragraph}{4}{\p@rind{\@startsection{paragraph}{4}{\p@rind}
{\p@rbeforeskip}{\p@rafterskip}{\p@rstyl@}}
\def\@startsection{subparagraph}{4}{\sp@rind{\@startsection{subparagraph}{4}{\sp@rind}
{\sp@rbeforeskip}{\sp@rafterskip}{\sp@rstyl@}}
\def\statementabove#1{\def\th@bove{#1}}
\def\statementstyle#1{\def\thstyl@{#1}}
\def\statementbelow#1{\def\thb@low{#1}}
\def\let\thind@nt\relax{\let\thind@nt\relax}
\def\let\thind@nt\indent{\let\thind@nt\indent}
\def\statementintrostyle#1{\def\thintr@style{#1}}
\def\statementintrodot#1{\def\thintr@dot{#1}}
\def\statementintrosep#1{\def\thintr@sep{#1}}
\def\statementintrobrackets#1#2{\def\thintr@left{#1}\def\thintr@right{#2}}
\statementabove{\vskip\medskipamount}
\statementstyle{\sl}
\statementbelow{\vskip\medskipamount}
\let\thind@nt\indent
\statementintrostyle{\normalshape\texbf\boldmath}
\statementintrodot{.}
\statementintrosep{\kern1.25ex}
\statementintrobrackets{(}{)}
\def\@thskip{\dimen100\lastskip\vskip-\dimen100%
\th@bove\dimen101\lastskip\vskip-\dimen101%
\ifdim\dimen100>\dimen101\else\dimen100\dimen101\fi\vskip\dimen100\vskip0pt}
\long\def\@@newtheorem#1#2#3{%
\newenvironment{#3}%
{\def\object@type{#3}\par\@thskip%
\@ifnextchar[{\@enva{#3}{\thstyl@#1{#2}}}{\@envb{#3}{\thstyl@#1{#2}}}}%
{\end{#3@}}%
\@ifnextchar[{\@nothm{#3}}{\@nnthm{#3}}}
\def\@nothm#1[#2]#3{%
\@ifundefined{c@#2}{\@latexerr{No theorem environment `#2' defined}\@eha}%
{\expandafter\@ifdefinable\csname #1@\endcsname
{\global\@namedef{the#1}{\@nameuse{the#2}}%
\global\@namedef{c@#1}{\@nameuse{c@#2}}
\global\@namedef{p@#1}{\@nameuse{p@#2}}
\global\@namedef{#1@}{\@nnnthm{#2}{#3}}%
\global\@namedef{end#1@}{\@endtheorem}}}}
\def\@nnnthm#1#2{\refstepcounter
{#1}\@ifnextchar[{\@ynnnthm{#1}{#2}}{\@xnnnthm{#1}{#2}}}
\def\@xnnnthm#1#2{\@begintheorem{#2}{\csname the#1\endcsname}\ignorespaces}
\def\@ynnnthm#1#2[#3]{\@opargbegintheorem{#2}{\csname the#1\endcsname}{#3}\ignorespaces}
\def\renewtheorem{\@ifnextchar[{\@renewtheorem}{\@renewtheorem[{}{}]}
\long\def\@renewtheorem[#1]{\@@renewtheorem#1}
\long\def\@@renewtheorem#1#2#3{%
\expandafter\let\csname#3@\endcsname\undefined
\renewenvironment{#3}%
{\def\object@type{#3}\par\@thskip%
\@ifnextchar[{\@enva{#3}{\thstyl@#1{#2}}}{\@envb{#3}{\thstyl@#1{#2}}}}%
{\end{#3@}}%
\@ifnextchar[{\@nothm{#3}}{\@nnthm{#3}}}
\def\@begintheorem#1#2{\@opargbegintheorem{#1}{#2}{}}
\def\@opargbegintheorem#1#2#3{%
\edef\@tempx{#1}%
\expandafter\let\expandafter\@tempy#
\def\@tempz{#3}%
\mytrivlist\@ifnextchar[{\item@}{\item@@}[\thind@nt\hskip\labelsep%
{\thintr@style%
#1\ifx\@tempx\@empty\else\ifx\@tempy\relax\else\kern1ex\fi\fi#2%
\ifx\@tempz\@empty%
\ifx\@tempx\@empty\ifx\@tempy\relax%
\else\thintr@dot\thintr@sep\fi\else\thintr@dot\thintr@sep\fi%
\else%
\ifx\@tempx\@empty\ifx\@tempy\relax\else\kern1ex\fi\else\kern1ex\fi%
\thintr@left{#3}\thintr@right\thintr@dot\thintr@sep\fi}%
\hskip-\labelsep]%
\ifautolabel\label*{#3}\fi}
\def\@endtheorem{\strut\endtrivlist\thb@low}
\defProof{Proof}
\def\proofabove#1{\def\pf@bove{#1}}
\def\proofstyle#1{\def\pfstyl@{#1}}
\def\proofbelow#1{\def\pfb@low{#1}}
\def\let\pfind@nt\relax{\let\pfind@nt\relax}
\def\let\pfind@nt\indent{\let\pfind@nt\indent}
\def\proofintrostyle#1{\def\pfintr@style{#1}}
\def\proofintrodot#1{\def\pfintr@dot{#1}}
\def\proofintrosep#1{\def\pfintr@sep{#1}}
\def\proofintrobrackets#1#2{\def\pfintr@left{#1}\def\pfintr@right{#2}}
\proofabove{\vskip\medskipamount}
\proofstyle{}
\proofbelow{\vskip\medskipamount}
\let\pfind@nt\indent
\proofintrostyle{\sl}
\proofintrodot{.}
\proofintrosep{\kern1.25ex}
\proofintrobrackets{of\kern1ex}{}
\def\@pfskip{\dimen100\lastskip\vskip-\dimen100%
\pf@bove\dimen101\lastskip\vskip-\dimen101%
\ifdim\dimen100>\dimen101\else\dimen100\dimen101\fi\vskip\dimen100\vskip0pt}
\renewenvironment{proof}%
{\@pfskip\mytrivlist\@ifnextchar[{\item@}{\item@@}[\pfind@nt]\@ifnextchar[{\pro@f}{\pro@f[\prooftag]}}
{\ifvoid\provedbox\else\hproved\fi\endtrivlist\pfb@low}
\def\pro@f[#1]{\setbox\provedbox\hbox{\provedboxcontents{#1}}\proofintro{#1}}
\def\proofintro#1{\expandafter\def\expandafter\@tempa\expandafter{#1}%
{\pfintr@style{Proof\ifx\@tempa\empty\else\kern1ex\pfintr@left{#1}%
\pfintr@right\fi}\pfintr@dot\pfintr@sep}\pfstyl@\ignorespaces}
\def\provedmark#1{\def\prm@rk{#1}}
\def\provedsep#1{\def\prs@p{#1}}
\provedmark{$\square$}
\provedsep{\kern1.25ex}
\def\def\prb@x##1{\fbox{\small##1}}{\def\prb@x##1{\fbox{\small##1}}}
\def\def\prb@x##1{\prm@rk}{\def\prb@x##1{\prm@rk}}
\def\let\prhf@l\hfill{\let\prhf@l\hfill}
\def\let\prhf@l\relax{\let\prhf@l\relax}
\def\prb@x##1{\prm@rk}
\let\prhf@l\relax
\def\provedboxcontents#1{\expandafter\def\expandafter\@tempa\expandafter{#1}%
\ifx\@tempa\empty\prm@rk\else\prb@x{#1}\fi}
\def\ifmmode\eqno{\box\provedbox}\else\hproved\fi{\ifmmode\eqno{\box\provedbox}\else\hproved\fi}
\def\hproved{\unskip\nobreak\prhf@l\penalty50\prs@p\hbox{}\nobreak\prhf@l
\box\provedbox{\finalhyphendemerits=0\par}}
\def\captionstyle#1{\def\c@ptstyl@{#1}}
\def\captionintrostyle#1{\def\c@pintr@style{#1}}
\def\captionintrodot#1{\def\c@pintr@dot{#1}}
\def\captionintrosep#1{\def\c@pintr@sep{#1}}
\captionstyle{\small\sf}
\captionintrostyle{\texbf\boldmath}
\captionintrodot{.}
\captionintrosep{\hskip1.25ex}
\long\def\@makecaption#1#2{%
\vskip\captionskip
\setbox\@tempboxa\hbox{%
\ifproofing\@ifundefined{the@label}{}
{\hbox to 0pt{\vbox to 0pt{\vss\hbox{\tiny\the@label}\vskip\bigskipamount}\hss}}\fi
\c@ptstyl@{\c@pintr@style #1\c@pintr@dot}\ignorespaces #2}%
\@captionwidth=\hsize \advance\@captionwidth-2\@captionmargin
\ifdim \wd\@tempboxa >\@captionwidth {%
\rightskip=\@captionmargin\leftskip=\@captionmargin
\unhbox\@tempboxa\par}%
\else
\hbox to\hsize{\hfil\box\@tempboxa\hfil}%
\fi}
\def\end@Float#1{%
\expandafter\caption\expandafter[\the@title]{%
{\c@pintr@style%
\ifx\the@caption\empty\ifx\the@title\empty
\else\c@pintr@sep\fi\else\c@pintr@sep\fi
\the@title\ifx\the@caption\empty%
\expandafter\label\expandafter*\expandafter{\the@label}%
\else\ifx\the@title\empty%
\expandafter\label\expandafter*\expandafter{\the@label}%
\else\c@pintr@dot\c@pintr@sep%
\expandafter\label\expandafter*\expandafter{\the@label}\fi\fi}%
\ignorespaces\the@caption}%
\end{#1}}
\renewenvironment{Figure}{\@ifnextchar[%
{\@myFloat{figure}}{\@myFloat{figure}[htbp]}}{\end@Float{figure}}
\newenvironment{Diagram}{\@ifnextchar[%
{\@myFloat{diagram}}{\@myFloat{diagram}[htbp]}}{\end{diagram}}
\def\@myFloat#1[#2]#3{%
\begin{#1}[#2]\def\the@label{#3}}
\def\showfigurestrue\fig{\let\fig@\fig@@@\fig}
\def\fig#1{\@ifnextchar[{\@fig{#1}}{\@fig{#1}[0pt]}}
\def\@fig#1[#2]#3{\@ifnextchar[{\@@fig{#1}[#2]{#3}}{\@@fig{#1}[#2]{#3}[0pt]}}
\def\@@fig#1[#2]#3[#4]#5#6{%
\def\the@title{#5}\def\the@caption{#6}\centerline{\fig@{#1}{#2}{#3}}\vskip#4}
\def\fig@@#1#2#3{\leavevmode{\figstyl@\vrule width 0pt height 1.8ex%
\smash{\framebox{\strut\def\@temp{#1}\ifx\@temp\@empty{ #3 }\else{ #1 }\fi}}}}
\def\fig@@@#1#2#3{\leavevmode\kern#2\epsfbox{#3}}
\def\figstyle#1{\def\figstyl@{#1}}
\figstyle{\family{cmr}\shape{n}\series{m}\selectfont\normalsize}
\def\oldFigure{%
\renewenvironment{Figure}{\@ifnextchar[%
{\@Float{figure}}{\@Float{figure}[htbp]}}{\end@Float{figure}}
\let\showfigurestrue\fig\@ldshowfig \let\fig\@ldfig
\let\let\fig@\fig@@@\@ldshowfigurestrue
\let\let\fig@\fig@@\@ldshowfiguresfalse
\let\fig@\fig@@}
\def\@ldshowfig#1#2{\epsfbox{#2}}
\def\@ldfig@#1#2{\leavevmode{\framebox{\figstyl@\strut{ #1 }}}}
\def\@ldshowfigurestrue{\let\fig\@ldshowfig}
\def\@ldshowfiguresfalse{\let\fig\@ldfig@}
\newcounter{diagram}
\let\thediagram\theequation
\def\ftype@diagram{2}
\def\ext@diagram{lod}
\def\@float{diagram}{\@float{diagram}}
\let\enddiagram\end@float
\newif\if@diagnum
\def\let\diag@\diag@@@\diag{\let\diag@\diag@@@\diag}
\def\diag#1{\@ifnextchar[{\@diag{#1}}{\@diag{#1}[0pt]}}
\def\@diag#1[#2]#3{\@ifnextchar[{\@@diag{#1}[#2]{#3}}{\@@diag{#1}[#2]{#3}[0pt]}}
\def\@@diag#1[#2]#3[#4]#5{
\def\the@tag{#5}\@eqnswtrue%
\centerline{\setbox0\hbox{\diag@{#1}{#2}{#3}}
\dimen0 -0.5\wd0\dimen1 0.5\ht0\box0%
\advance\dimen0 0.5\hsize\advance\dimen0 -\rightskip\advance\dimen1 #4%
\let\@currentlabel\the@tag%
\setbox0\hbox to 0pt{\hss\family{cmr}\shape{n}\series{m}\selectfont(\the@tag)}%
\ifx\the@tag\@empty\refstepcounter{equation}\let\@currentlabel\theequation%
\setbox0\hbox to 0pt{\hss\family{cmr}\shape{n}\series{m}\selectfont(\thediagram)}\fi%
\if@eqnsw\else\let\@currentlabel\relax\setbox0\hbox to 0pt{}\fi%
\advance\dimen1 -0.5\ht0%
\@ifnextchar[{\@put}{\@@rput[\z@,\z@][r]}[\dimen0,\dimen1][l]{%
\box0\expandafter\label\expandafter*\expandafter{\the@label}\kern0.15em}}}
\def\diag@@#1#2#3{\leavevmode{\diagstyl@\vrule width 0pt height 1.8ex%
\smash{\framebox{\strut\def\@temp{#1}\ifx\@temp\@empty{ #3 }\else{ #1 }\fi}}}}
\def\diag@@@#1#2#3{\leavevmode\kern#2\epsfbox{#3}}
\def\diagstyle#1{\def\diagstyl@{#1}}
\diagstyle{\family{cmr}\shape{n}\series{m}\selectfont\normalsize}
\def\let\fig@\fig@@{\let\fig@\fig@@}
\def\let\fig@\fig@@@{\let\fig@\fig@@@}
\def\let\diag@\diag@@{\let\diag@\diag@@}
\def\let\diag@\diag@@@{\let\diag@\diag@@@}
\def\n@number{\@eqnswfalse\let\@currentlabel\relax\let\the@tag\relax}
\def\equation{$
\@eqnswtrue\def\object@type{equation}\let\nonumber\n@number%
\advance\c@equation1\edef\@currentlabel{\theequation}\advance\c@equation-1%
\def\the@tag{\refstepcounter{equation}\eqno\hbox{\@eqnnum}}}
\def\tag#1{\edef\@currentlabel{#1}\def\the@tag{\eqno\hbox{\reset@font\ifmmode\else\protect\prm\fi(#1)}}}
\def\endequation{\the@tag$
\global\@ignoretrue}
\dimen200\topsep
\dimen201\partopsep
\topsep0pt
\partopsep0pt
\let\ifmmode\else\protect\pit\fi@m\@ifnextchar[{\item@}{\item@@}
\def\@ifnextchar[{\item@}{\item@@}{\@ifnextchar[{\@ifnextchar[{\item@}{\item@@}@}{\@ifnextchar[{\item@}{\item@@}@@}}
\def\@ifnextchar[{\item@}{\item@@}@[#1]{\ifmmode\else\protect\pit\fi@m[#1]\vskip-\lastskip\vskip\itemsep}
\def\@ifnextchar[{\item@}{\item@@}@@{\ifmmode\else\protect\pit\fi@m\vskip-\lastskip\vskip\itemsep}
\def\s@titemsep{\@ifnextchar[{\s@@titemsep}{\relax}}
\def\s@@titemsep[#1]{\itemsep#1}
\let\@itemize\itemize
\let\@enditemize\enditemize
\renewenvironment{itemize}
{\@itemize\itemsep3pt\parsep0pt\topsep0pt\partopsep0pt\s@titemsep}
{\@enditemize\vskip-\lastskip\vskip\itemsep}
\let\@enumerate\enumerate
\let\@endenumerate\endenumerate
\renewenvironment{enumerate}
{\@enumerate\itemsep3pt\parsep0pt\topsep0pt\partopsep0pt\s@titemsep}
{\@endenumerate\vskip-\lastskip\vskip\itemsep}
\let\@description\description
\let\@enddescription\enddescription
\renewenvironment{description}
{\@description\itemsep3pt\parsep0pt\topsep0pt\partopsep0pt\s@titemsep}
{\@enddescription\vskip-\lastskip\vskip\itemsep}
\topsep\dimen200
\partopsep\dimen201
\@definecounter{bibenumi}
\def\thebibliography#1{%
\@startsection{section}{1}{\s@ctind*{\refname}\vskip-\lastskip%
\list{[\arabic{bibenumi}]}{\topsep0pt\settowidth\labelwidth{[#1]}%
\leftmargin\labelwidth\advance\leftmargin\labelsep\usecounter{bibenumi}}%
\def\newblock{\hskip .11em plus .33em minus .07em}%
\sloppy\clubpenalty4000\widowpenalty4000\sfcode`\.=1000\relax}
\makeatother
\parsep0pt
\topsep0pt
\itemsep0pt
\partopsep0pt
\makeatletter
\c@totalnumber8
\c@topnumber8
\c@bottomnumber8
\def0{0}
\def1{1}
\def1{1}
\def0{0}
\textfloatsep\floatsep
\makeatother
\abovedisplayskip\smallskipamount
\belowdisplayskip\smallskipamount
\abovedisplayshortskip\smallskipamount
\belowdisplayshortskip\smallskipamount
\frenchspacing
\proofingfalse
\autolabelfalse
\let\fig@\fig@@@
\let\diag@\diag@@@
\newtheorem{stat}{#1}\begin{stat}}{\end{stat}} \unnumbered{stat}
\newenvironment{statement}[1]{\def#1}\begin{stat}}{\end{stat}{#1}\begin{stat}}{\end{stat}}
\newtheorem{nstat}{#1}\begin{nstat}}{\end{nstat}}[section]
\newenvironment{numberedstatement}[1]{\def#1}\begin{nstat}}{\end{nstat}{#1}\begin{nstat}}{\end{nstat}}
\newtheorem{axiom}[nstat]{Axiom}
\newtheorem{postulate}[nstat]{Postulate}
\newtheorem{definition}[nstat]{Definition}
\newtheorem{lemma}[nstat]{Lemma}
\newtheorem{proposition}[nstat]{Proposition}
\newtheorem{theorem}[nstat]{Theorem}
\newtheorem{corollary}[nstat]{Corollary}
\newtheorem{conjecture}[nstat]{Conjecture}
\newtheorem{problem}[nstat]{Problem}
\newtheorem{question}[nstat]{Question}
\newtheorem{exercise}[nstat]{Exercise}
\newtheorem{example}[nstat]{Example}
\newtheorem{remark}[nstat]{Remark}
\let\ns\normalshape
\papersize{216truemm}{279truemm}
\@ifnextchar[{\@margins}{\@margins[\z@]}{33truemm}{20.5truemm}
\advance\voffset-5mm
\advance\hoffset7.5truemm
\advance\footskip2.5truemm
\if@twoside\let\n@xt\h@dlin@\else\let\n@xt\h@@dlin@\fi\n@xt{\hfill}
\if@twoside\let\n@xt\f@tlin@\else\let\n@xt\f@@tlin@\fi\n@xt{\small\hfill--\kern1ex\thepage\kern1ex--\hfill}
\frenchspacing
\flushbottom
\makeatletter
\c@totalnumber8
\c@topnumber8
\c@bottomnumber8
\def0{0}
\def1{1}
\def1{1}
\def0{0}
\makeatother
\textfloatsep\floatsep
\abovedisplayskip\smallskipamount
\belowdisplayskip\smallskipamount
\abovedisplayshortskip\smallskipamount
\belowdisplayshortskip\smallskipamount
\lineskiplimit-6pt
\arraycolsep2pt
\let\fig@\fig@@@
\sectionbeforeskip{1.5\bigskipamount}
\sectionstyle{\centering\normalsize\sc}
\sectionafterskip{\bigskipamount}
\@startsection{section}{1}{\s@ctind@ftindtrue
\subsectionbeforeskip{\bigskipamount}
\subsectionstyle{\normalsize\texbf\boldmath}
\subsectionafterskip{.5\bigskipamount}
\@startsection{subsection}{2}{\ss@ctind@ftindtrue
\paragraphbeforeskip{\medskipamount}
\paragraphstyle{\texbf\boldmath\sl}
\paragraphafterskip{1.25ex}
\def\p@rind{\parindent}
\paragraphdot{.}
\newtheorem{gnstat}{\gnstatname}[section]
\statementintrostyle{\texbf\boldmath}
\renewtheorem{lemma}[gnstat]{Lemma}
\renewtheorem{proposition}[gnstat]{Proposition}
\renewtheorem{theorem}[gnstat]{Theorem}
\renewtheorem{corollary}[gnstat]{Corollary}
\renewtheorem{problem}[gnstat]{Problem}
\renewtheorem[{\ns}{}]{definition}[gnstat]{Definition}
\renewtheorem[{\ns}{}]{remark}[gnstat]{Remark}
\captionstyle{\small\ns}
\captionintrostyle{\sc}
\def~{\kern0.33em}
\let\mycal\cal
\def\cal#1{{\mycal #1}}
\let\mymathrm\mathrm
\def\mathrm#1{{\mymathrm #1}}
\let\texbf\texbf\boldmath
\def\texbf\boldmath{\texbf\boldmath}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left.eps}}}}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right.eps}}}}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both.eps}}}}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left-long.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left-long.eps}}}}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}}
\def\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both-long.eps}}}{\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both-long.eps}}}}
\let\rightmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right.eps}}}
\let\leftmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left.eps}}}
\let\leftrightmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both.eps}}}
\let\longrightmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}
\let\longleftmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-left-long.eps}}}
\let\longleftrightmove\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-both-long.eps}}}
\def{\text{\raise.21ex\hbox{$\not$}}\mkern.15mu\mathrm{O}\mkern.15mu}{{\text{\raise.21ex\hbox{$\not$}}\mkern.15mu\mathrm{O}\mkern.15mu}}
\let\emptyset{\text{\raise.21ex\hbox{$\not$}}\mkern.15mu\mathrm{O}\mkern.15mu}
\let\0\emptyset
\let\theta\textheta
\let\texbar\bar
\let\texhat\hat
\let\textilde\tilde
\let\bar\widebar
\let\hat\widehat
\let\tilde\widetilde
\let\longto\longrightarrow
\let\longleadsto\mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}
\newcommand{\cal A}{\cal A}
\newcommand{\cal D}{\cal D}
\newcommand{\cal E}{\cal E}
\newcommand{\cal G}{\cal G}
\newcommand{\cal I}{\cal I}
\newcommand{\cal M}{\cal M}
\newcommand{\cal R}{\cal R}
\renewcommand{\S}{\cal S}
\newcommand{\cal T}{\cal T}
\newcommand{\mathop{\mathrm{id}}\nolimits}{\mathop{\mathrm{id}}\nolimits}
\newcommand{\mathop{\mathrm{Cl}}\nolimits}{\mathop{\mathrm{Cl}}\nolimits}
\newcommand{\mathop{\mathrm{Int}}\nolimits}{\mathop{\mathrm{Int}}\nolimits}
\newcommand{\mathop{\mathrm{Stab}}\nolimits}{\mathop{\mathrm{Stab}}\nolimits}
\newcommand{\mathop{\mathrm{Orb}}\nolimits}{\mathop{\mathrm{Orb}}\nolimits}
\newcommand{\mathop{\mathrm{len}}\nolimits}{\mathop{\mathrm{len}}\nolimits}
\newcommand{\mathop{\mathrm{ord}}\nolimits}{\mathop{\mathrm{ord}}\nolimits}
\newcommand{\mathop{\mathrm{LCM}}\nolimits}{\mathop{\mathrm{LCM}}\nolimits}
\begin{document}
\title{\large\texbf\boldmath
AUTOMORPHISMS OF TRIVALENT GRAPHS
\label{Version 1.0 / \today}}
\author{\normalsize\sc Silvia Benvenuti\\
\normalsize\sl Scuola di Scienze e Tecnologie\\
\normalsize\sl Universit\`a di Camerino -- Italia\\
\small\tt <EMAIL>
\and
\normalsize\sc Riccardo Piergallini\\
\normalsize\sl Scuola di Scienze e Tecnologie\\
\normalsize\sl Universit\`a di Camerino -- Italia\\
\small\tt <EMAIL>}
\date{}
\vglue1.2cm
\maketitle
\vskip\bigskipamount
\begin{abstract}
\baselineskip13.5pt
\vskip\medskipamount
\noindent
Let $\cal G_{g,b}$ be the set of all uni/trivalent graphs representing the combinatorial
structures of pant decompositions of the oriented surface $\Sigma_{g,b}$ of genus $g$ with
$b$ boundary components. We describe the set $\cal A_{g,b}$ of all automorphisms of graphs in
$\cal G_{g,b}$ showing that, up to suitable moves changing the graph within $\cal G_{g,b}$, any
such automorphism can be reduced to elementary switches of adjacent edges.
\vskip\medskipamount\vskip\smallskipamount\noindent
{\sl Keywords}\/: uni/trivalent graph, $F$-move, pant decomposition complex.
\vskip\medskipamount\noindent
{\sl AMS Classification}\/: Primary 20B25; Secondary 57M50, 05C25.
\end{abstract}
\maketitle
\vskip\bigskipamount
\@startsection{section}{1}{\s@ctind*{Introduction}
Let $\cal G_{g,b}$ be the set of all connected non-oriented uni/trivalent graphs $\Gamma$,
with first Betti number $\beta_1(\Gamma) = g \geq 0$, at least one trivalent vertex and $b
\geq 0$ univalent vertices. It is immediate to see that such a graph $\Gamma$ exists if
and only if $(g,b) \neq (0,0),(0,1),(0,2),(1,0)$, and in this case $\Gamma$ has $2g-2+b$
trivalent vertices and $3g-3+2b$ edges. In the following, the pair $(g,b)$ will be always
assumed to satisfy that restriction.
Given any $\Gamma \in \cal G_{g,b}$, we denote by $\cal A(\Gamma)$ the group of all the
automorphisms of $\Gamma$ as a non-oriented graph. In this paper we are interested in the
structure of the set $\cal A_{g,b} = \cup_{\Gamma \in \cal G_{g,b}}\cal A(\Gamma)$ of all the
automorphisms of graphs in $\cal G_{g,b}$. More precisely, we will prove that, up to certain
$F$-moves changing the graph within $\cal G_{g,b}$, any such automorphism can be reduced to
elementary switches of adjacent edges.
The drive for this paper comes from a problem we met while working on pant decompositions
of surfaces in \cite{BP08}. Namely, in that paper we build up an infinite simply connected
2-dimensional complex $\cal R_{g,b}$, codifying all the pant decompositions on the connected
compact oriented surface $\Sigma_{g,b}$ of genus $g$ with $b$ boundary components, as well
as all the moves relating them and all the relations between those moves. The construction
is subdivided into two independent steps.
First a finite simply connected 2-dimensional complex $\S_{g,b}$ is described, which
codifies the {\sl combinatorial structures} of pant decompositions of $\Sigma_{g,b}$. The
combinatorial structure of a pant decomposition $D$ consists of the information concerning
only the incidence relations between pants, and it is encoded by the dual graph $\Gamma_D
\in \cal G_{g,b}$, with the trivalent vertices corresponding to the pants, the univalent
vertices corresponding to the boundary components, and the edges corresponding to the
cutting curves. Afterwards, by using a presentation of the mapping class group $\cal M_{g,b}$
of $\Sigma_{g,b}$, we get the desired complex $\cal R_{g,b}$.
Due to technical issues, the switch from $\S_{g,b}$ to $\cal R_{g,b}$ is not a ``direct''
one, but it requires instead a cumbersome intermediate step. This involves in the
construction of two additional complexes $\tilde{\S}_{g,b}$ and $\tilde{\cal R}_{g,b}$,
codifying respectively the {\sl decorated combinatorial structures} and the {\sl decorated
pant decompositions}. Here, a decoration of a pant decomposition $D$ is a numbering of its
cutting curves, and a decoration of the corresponding combinatorial structure is a
numbering of the edges of $\Gamma_D$.
This method allows us to sidestep a crucial problem in the direct transition from
$\S_{g,b}$ to $\cal R_{g,b}$, that is the study of the stabilization subgroup $\mathop{\mathrm{Stab}}\nolimits(D) \in
\cal M_{g,b}$ of any given pant decomposition $D$. In fact, the proof of the simple
connectedness of $\cal R_{g,b}$ given in \cite{BP08} would become less elaborate based on the
fact that the fiber $p^{-1}(\Gamma_D)$ of the natural projection $p: \cal R_{g,b} \to
\S_{g,b}$ is trivial at $\cal R_{g,b}$ on the level of the fundamental group. In other words,
we need to show that any loop based at $D$ and contained in $p^{-1}(\Gamma_D)$ is
contractible in $\cal R_{g,b}$. Now, such a loop corresponds to an element of $\mathop{\mathrm{Stab}}\nolimits(D)$, that
is a symmetry of $D$, which induces a (possibly trivial) combinatorial symmetry of
$\Gamma_D$. Actually, once the combinatorial symmetries of $\Gamma_D$ are known, one can
reconstruct those of $D$ in a straightforward way, by adding some easy topological
information.
At this point, it should be clear the motivation for the study carried out in the present
paper of the structure of $\cal A_{g,b}$, and in particular of the inclusion $\cal A(\Gamma)
\subset \cal A_{g,b}$ of the group of combinatorial symmetries $\cal A(\Gamma)$ of a given $\Gamma
\in \cal G_{g,b}$. In fact, the result we obtain here will enable us to give a much simpler
contruction of $\cal R_{g,b}$ than in \cite{BP08}, and a more direct proof of its simple
connectedness starting from $\S_{g,b}$ without involving decorations.
However, the study of the automorphisms of uni/trivalent graphs has its own interest
independently from that specific application (see the classical papers \cite{Tu48,Tu59}
and \cite{Go80}, and some more recent ones, as for instance \cite{CD02,Ca04,FK06,FK07}).
Hence, in this paper we focus on the structure of $\cal A_{g,b}$, while the new construction
of $\cal R_{g,b}$ mentioned above will be described in details in a forthcoming paper
\cite{BP12}.
The paper is structured as follows. In Section~\ref{main/sec} we give the basic
definitions, the statement of the main theorem, and the reduction of its proof to special
cases. Then, after having established some preliminary results in
Section~\ref{lemmas/sec}, we devote the subsequent Sections~\ref{order-pm/sec},
\ref{order-3m/sec}, \ref{order-2m/sec} to those special cases.
\@startsection{section}{1}{\s@ctind{Definitions and main theorem%
\label{main/sec}}
For a graph $\Gamma \in \cal G_{g,b}$, we call a {\sl free end} of $\Gamma$ any univalent
vertex, a {\sl terminal edge} of $\Gamma$ any of the $b$ edges connecting a free end to a
trivalent vertex, and an {\sl internal edge} of $\Gamma$ any of the $3g - 3 + b$ edges
connecting two (possibly coinciding) trivalent vertices.
First of all, we introduce $F$-moves on uni/trivalent graphs. These are well-known moves
that change the graph structure by acting on the internal edges. It is a folklore result
that $F$-moves, even in the most restrictive form given in
Definition~\ref{Fmove-edge/def}, suffice to relate any two graphs in $\cal G_{g,b}$ up to
graph isomorphisms (cf. \cite{HT80}). However, we need to consider an invariant version of
$F$-moves, in order to relate graph automorphisms. This makes things more involved.
\begin{definition}\label{Fmove-edge/def}
Let $\Gamma \in \Gamma_{g,b}$ and $e \in \Gamma$ be an internal edge with distinct ends.
Then, we call {\sl elementary (edge) $F$-move}, the modification $F_{e,e'}: \Gamma
\rightmove \Gamma'$ that makes $\Gamma$ into $\Gamma' \in \cal G_{g,b}$, by replacing $e$ with
a new internal edge $e'$ as in Figure~\ref{move-edge/fig}, while leaving the rest of the
graph unchanged. Clearly, the inverse modification $F_{e',e}: \Gamma' \rightmove \Gamma$
is an elementary (edge) $F$-move as well.
\end{definition}
\begin{Figure}[htb]{move-edge/fig}
\fig{}{move-edge.eps}[-3pt]
{}{An edge $F$-move $F_{e,e'}$.}
\end{Figure}
According to the literature, here the letter $F$ stands for ``fusion'' and refers to the
fact that the $F$-move $F_{e,e'}$ can be thought as the contraction of the edge $e$ in
$\Gamma$, and the consequent fusion of its ends, followed by the inverse of a similar
contraction of the edge $e'$ in $\Gamma'$. The intermediate graph has a quadrivalent
vertex in place of the two trivalent ends of the contracted edges, as indicated under
the arrow.
We warn the reader that, since the graphs in $\cal G_{g,b}$ are abstract ones, properties
and constraints depending on the planar representation (of portions) of them in the
figures do not have any significance.
In particular, the coupling $(e_1 e_2)(e_3 e_4)$ of the (possibly not all distinct) edges
$e_1, \dots, e_4$ determined by the two ends of $e$ in $\Gamma$, which in
Figure~\ref{move-edge/fig} is replaced by the coupling $(e_1 e_4)(e_2 e_3)$ determined by
the two ends of $e'$ in $\Gamma'$, could be replaced by the coupling $(e_1 e_3)(e_2 e_4)$
as well. In other words, there are exactly two possible ways to perform the $F$-move
$F_{e,e'}$ at the given edge $e$ of $\Gamma$, corresponding to the two possible changes of
coupling, depending on the structure of $\Gamma'$ at the edge $e'$:
$$(e_1 e_2)(e_3 e_4) \mathrel{\buildrel \textstyle F_{e,e'} \over \mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}}
(e_1 e_4)(e_2 e_3)
\quad \text{and} \quad
(e_1 e_2)(e_3 e_4) \mathrel{\buildrel \textstyle F_{e,e'} \over \mathrel{\raise0.8pt\hbox{\epsfbox{wavearrow-right-long.eps}}}}
(e_1 e_3)(e_2 e_4).$$
From a different perspective, the elementary $F$-move described in
Figure~\ref{move-edge/fig}, can be interpreted as the replacement of a uni/trivalent
subtree $T \subset \Gamma$ with a different uni/trivalent subtree $T' \subset \Gamma'$
having the same univalent vertices. Namely, $T$ consists of the edges $e_1, \dots, e_4$
and $e$, while $T'$ consists of the edges $e_1, \dots, e_4$ and $e'$. This suggests the
following generalization of the notion of elementary $F$-move.
\begin{definition}\label{Fmove-tree/def}
Let $\Gamma \in \cal G_{g,b}$ and $T \subset G$ be a uni/trivalent subtree with $m \geq 4$
free ends. Then, we call {\sl elementary (tree) $F$-move}, the modification $F_{T,T'}:
\Gamma \rightmove \Gamma'$ that makes $\Gamma$ into $\Gamma' \in \cal G_{g,b}$, by replacing
$T$ with a different uni/trivalent tree $T' \subset \Gamma'$ having the same free ends of
$T$, while leaving the rest of the graph unchanged (see Figure~\ref{move-tree/fig} for an
example with $m = 5$). Also in this case, the inverse modification $F_{T',T}: \Gamma'
\rightmove \Gamma$ is an elementary (tree) $F$-move as well.
\end{definition}
\begin{Figure}[htb]{move-tree/fig}
\fig{}{move-tree.eps}[-3pt]
{}{A tree $F$-move $F_{T,T'}$.}
\end{Figure}
We note that the $F$-move $F_{T,T'}$ is again a kind of fusion move. In fact, it can be
thought as the contraction of the subtree consisting of all the internal edges of $T
\subset \Gamma$ to a single $m$-valent vertex, and the consequent fusion of all the
trivalent vertices of $T$, followed by the inverse of a similar contraction for $T'
\subset \Gamma'$.
If the subtree $T \subset \Gamma$ has the vertices $v_1, \dots, v_m$ as its free ends,
then all the possible ways to perform the $F$-moves $F_{T,T'}$ on $\Gamma$ corresponds to
all the complete iterated couplings of the (unordered) set $\{v_1, \dots ,v_m\}$ induced
by $T'$, except the one induced by $T$. In particular, according to the above discussion,
we have 2 ways for $m = 4$.
Actually, any tree $F$-move $F_{T,T'}$ can be realized by a suitable sequence of edge
$F$-moves performed on the edges of $T$ (and of the corresponding subtrees in the
resulting graphs). Yet, it makes sense to consider $F_{T,T'}$ as a unique move, and we
will see shortly why (cf. discussion about the $F$-move in Figure~\ref{move-inv/fig}
below).
\begin{definition}\label{Fmove/def}
Let $\Gamma \in \cal G_{g,b}$ and $\cal T = \{T_1, \dots, T_k\}$ be a family of uni/trivalent
subtrees of $\Gamma$ with pairwise disjoint interiors (that is, two of them possibly share
only some common free ends). Assuming that each $T_i$ has at least 4 free ends, we call
{\sl $F$-move} any modification $F_{\cal T,\cal T'}: \Gamma \rightmove \Gamma'$ given by the
simultaneous application of certain elementary moves $F_{T_1,T'_1}, \dots, F_{T_k,T'_k}$
to $\Gamma$. Of course, for $k = 1$ we have an elementary $F$-move.
\end{definition}
For any graph $\Gamma \in \cal G_{g,b}$ we denote by $\cal A(\Gamma)$ the group of all
the automorphisms of $\Gamma$ as a non-oriented graph, including those that permute the
free ends. Moreover, we denote by $\cal A_{g,b} = \cup_{\Gamma \in \cal G_{g,b}}\cal A(\Gamma)$ the
set of all the automorphisms of graphs in $\cal G_{g,b}$.
\vskip\medskipamount
Now, assume we are given an automorphism $\phi \in \cal A(\Gamma)$. If $F_{\cal T,\cal T'}: \Gamma
\rightmove \Gamma'$ is an $F$-move with $\cal T = \{T_1, \dots, T_k\}$ a $\phi$-invariant
family of subtrees of $\Gamma$, then the restriction of $\phi$ to $\Gamma - \cup_{i} \mathop{\mathrm{Int}}\nolimits
T_i$ induces a graph automorphism $\psi$ of $\Gamma' - \cup_i\mathop{\mathrm{Int}}\nolimits T'_i$, under\break the
canonical isomorphism $\Gamma - \cup_i \mathop{\mathrm{Int}}\nolimits T_i \cong \Gamma' - \cup_i \mathop{\mathrm{Int}}\nolimits T'_i$ of
graphs given by $F_{\cal T,\cal T'}$. In general, such $\psi$ does not extend to an automorphism
$\phi' \in \cal A(\Gamma')$, but if it does then the extension $\phi': \Gamma' \to \Gamma'$
can be easily seen to be unique, by taking into account that $\Gamma - \cup_i \mathop{\mathrm{Int}}\nolimits T_i$
contains all the free ends of all the subtrees $T_i$. When $\phi'$ exists, we say that it
is induced by $\phi$ through $F_{\cal T,\cal T'}$.
\begin{definition}\label{Fmove-inv/def}
Given an automorphism $\phi: \Gamma \to \Gamma$, we say that an $F$-move $F_{\cal T,\cal T'}:
\Gamma \rightmove \Gamma'$ is {\sl $\phi$-invariant} if $\cal T$ is a $\phi$-invariant
family of subtrees and $\phi$ induces an automorphism $\phi': \Gamma' \to \Gamma'$ through
$F_{\cal T,\cal T'}$ as discussed above. In this case, we write $F_{\cal T,\cal T'}: \phi \rightmove
\phi'$. Such relation between automorphisms is symmetric, in the sense that $F_{\cal T',\cal T}:
\phi' \rightmove \phi$ holds as well, being $F_{\cal T',\cal T}: \Gamma' \rightmove \Gamma$ a
$\phi'$-invariant $F$- move. But it is neither reflexive nor transitive. Then, we call
{\sl F-equivalence} the generated equivalence relation on the set $\cal A_{g,b}$.
\end{definition}
We observe that the elementary $F$-move in Figure~\ref{move-edge/fig} is $\phi$-invariant
with respect to any automorphism $\phi \in \cal A(\Gamma)$ that acts on the depicted
portion of $\Gamma$ as any planar symmetry (horizontal, vertical or central). On the
contrary, it is not $\phi$-invariant for a $\phi$ that switches $e_1$ and $e_2$ while
leaving $e_3$ and $e_4$ fixed.
\begin{Figure}[htb]{move-inv/fig}
\fig{}{move-inv.eps}[-3pt]
{}{A $\phi$-invariant $F$-move.}
\end{Figure}
Figure~\ref{move-inv/fig} shows a $\phi$-invariant elementary $F$-move $F_{T,T'}$, where
$\phi \in \cal A(\Gamma)$ is any automorphism that acts on $T$ as the simmetry with respect to
the vertical axis. As we said above, $F_{T,T'}$ could be realized by a sequence of
elementary edge $F$-moves, but this is not true if we insist that each single edge
$F$-move is $\phi$-invariant. Finally, notice that instead the $F$-move described in
Figure~\ref{move-tree/fig} is not $\phi$-invariant (for the same automorphism $\phi$).
\begin{remark}\label{slidings/rem}
In the following, we will essentially use only $\phi$-invariant $F$-moves $F_{\cal T,\cal T'}$
consisting of simultaneous elementary edge $F$-moves $F_{e,e'}$ or tree $F$-moves
$F_{T,T'}$ like the one in Figure~\ref{move-inv/fig}. We emphasize once again that the
$\phi$-invariance of $F$-moves $F_{\cal T,\cal T'}$ of does not imply the $\phi$-invariance of
each single elementary $F$-move, being these possibly permuted by $\phi$. In many cases,
it will be convenient to think of those elementary moves as the edges slidings shown in
Figure~\ref{move-3d/fig}. Here, the edges $f,f_1$ and $f_2$ are the original ones of
$\Gamma$, whose ends in the figure are slided to get their new positions in $\Gamma'$ as
indicated by the arrows, while the rest of the graph is fixed.
\end{remark}
\begin{Figure}[htb]{move-3d/fig}
\fig{}{move-3d.eps}[-3pt]
{}{$F$-moves as edge slidings.}
\end{Figure}
At this point, in order to state our main theorem, we are left to introduce the elementary
automorphisms, which will generate all the automorphisms in $\cal A_{g,b}$ up to
$F$-equivalence.
\begin{definition}\label{switches/def}
By an {\sl elementary automorphism} of a graph $\Gamma \in \cal G_{g,b}$ we mean any
automorphism $S_{e_1,e_2}: \Gamma \to \Gamma$ interchanging two adjacent edges $e_1$ and
$e_2$ of $\Gamma$, while fixing all the rest of the graph. We call such an automorphism
$S_{e_1,e_2} \in \cal A(\Gamma)$ a {\sl terminal switch} or an {\sl internal switch},
depending on the fact that $e_1$ and $e_2$ are terminal or internal edges (cf.
Figure~\ref{switches/fig}).
\end{definition}
\begin{Figure}[htb]{switches/fig}
\fig{}{switches.eps}[-3pt]
{}{A terminal switch (left side) and an internal switch (right side).}
\end{Figure}
\begin{remark}
An automorphism in $\Gamma \in \cal G_{g,b}$ is uniquely determined up to internal switches by
its action on the vertices of $\Gamma$. Moreover, it is uniquely determined up to
(terminal and internal) switches by its action on the trivalent vertices of $\Gamma$.
\end{remark}
To simplify our claims, we provide the next definition.
\begin{definition}\label{Egn/def}
For any $g$ and $b$, we denote by $\cal E_{g,b} \subset \cal A_{g,b}$ the smallest subset that
contains all the elementary automorphisms (internal and terminal switches) in $\cal A_{g,b}$
and is closed with respect to composition and $F$-equivalence of automorphisms.\hfill
\end{definition}
Then, our main result can be stated as follows.
\begin{theorem}\label{main/thm}
For any $g$ and $b$, we have $\cal E_{g,b} = \cal A_{n,b}$.
\end{theorem}
\begin{proof}
We have to show that any element $\phi \in \cal A_{g,b}$, that is any automorphism $\phi:
\Gamma \to \Gamma$ of a graph $\Gamma \in \cal G_{g,b}$, actually belongs to $\phi \in
\cal E_{g,b}$. By primary decomposition of the cyclic subgroup $\langle \phi \rangle \subset
\cal A(\Gamma)$, we can reduce ourselves to the special case when the order $\mathop{\mathrm{ord}}\nolimits(\phi)$, that
is the cardinality of $\langle\phi\rangle$, is a prime power. Then, the subcases when
$\mathop{\mathrm{ord}}\nolimits(\phi) = p^m$ with $p$ prime $> 3$ or $\mathop{\mathrm{ord}}\nolimits(\phi) = 3^m$ are reduced to the case
$\mathop{\mathrm{ord}}\nolimits(\phi) = 2$ in Sections~\ref{order-pm/sec} and \ref{order-3m/sec} respectively.
Finally, the case of $\mathop{\mathrm{ord}}\nolimits(\phi) = 2^m$ is proved in Section~\ref{order-2m/sec}.
\end{proof}
\@startsection{section}{1}{\s@ctind{Some preliminary results%
\label{lemmas/sec}}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a uni/trivalent graph $\Gamma$.
Two elements (vertices or edges) or subgraphs $x$ and $y$ of $\Gamma$ are said to be {\sl
$\phi$-equivalent} if $y = \phi^i(x)$ for some $i \geq 0$. In this case we write $x
\cong_\phi y$.
We denote by $\mathop{\mathrm{ord}}\nolimits(\phi)$ the order of $\phi$ in $\cal A(\Gamma)$ (as an automorphism of a
non-oriented graph), that is the cardinality of $\langle\phi\rangle \subset \cal A(\Gamma)$.
For a vertex $v$ of $\Gamma$, by the {\sl order of $v$ with respect to $\phi$}, we mean as
usual the cardinality $\mathop{\mathrm{ord}}\nolimits_\phi(v)$ of its $\phi$-orbit $\mathop{\mathrm{Orb}}\nolimits_\phi(v) = \{\phi^k(v) \,|\,
0 \leq k < \mathop{\mathrm{ord}}\nolimits(\phi)\}$. On the contrary, for an edge $e$ of $\Gamma$ with (possibly
coinciding) ends $v$ and $w$, we call {\sl order of $e$ with respect to $\phi$} the number
$\mathop{\mathrm{ord}}\nolimits_\phi(e) = \mathop{\mathrm{LCM}}\nolimits(\mathop{\mathrm{ord}}\nolimits_\phi(v), \mathop{\mathrm{ord}}\nolimits_\phi(w))$, which is equal to either the
cardinality of the $\phi$-orbit $\mathop{\mathrm{Orb}}\nolimits_\phi(e) = \{\phi^k(e) \,|\, 0 \leq k < \mathop{\mathrm{ord}}\nolimits(\phi)\}$
or its double (if $v \neq w$ and $(v,w) \cong_\phi (w,v)$ as ordered pairs). This amounts
to consider $e$ as an oriented edge, except the case when it is a loop.
\begin{lemma}\label{ordinv/thm}
If $\phi$ and $\phi'$ are $F$-equivalent in $\cal A_{g,b}$, then $\mathop{\mathrm{ord}}\nolimits(\phi) = \mathop{\mathrm{ord}}\nolimits(\phi')$.
\end{lemma}
\begin{proof}
It suffices to consider the case when $\phi$ and $\phi'$ are related by a single $F$-move
$F_{\cal T,\cal T'}: \phi \rightmove \phi'$, with $\cal T = \{T_1, \dots, T_k\}$ and $\cal T' = \{T'_1,
\dots, T'_k\}$. Then, $\mathop{\mathrm{ord}}\nolimits(\phi)$ coincides with the order of its restriction to $\Gamma
- \cup_i\, \mathop{\mathrm{Int}}\nolimits T_i$, by the uniqueness of extension to the subtrees $T_i$. Similarly,
$\mathop{\mathrm{ord}}\nolimits(\phi')$ coincides with the order of its restriction to $\Gamma' - \cup_i\, \mathop{\mathrm{Int}}\nolimits
T'_i$. Since those restrictions coincide under the canonical isomorphism $\Gamma - \cup_i
\mathop{\mathrm{Int}}\nolimits T_i \cong \Gamma' - \cup_i \mathop{\mathrm{Int}}\nolimits T'_i$, we have $\mathop{\mathrm{ord}}\nolimits(\phi) = \mathop{\mathrm{ord}}\nolimits(\phi')$.
\end{proof}
\begin{lemma}\label{orders/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$. If $v$ is a vertex of $\Gamma$ and $e$ is an edge of
$\Gamma$ having $v$ as an end, then we have one of the following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] $\mathop{\mathrm{ord}}\nolimits_\phi(e) =
\mathop{\mathrm{ord}}\nolimits_\phi(v)$;
\@ifnextchar[{\item@}{\item@@}[2)] $\mathop{\mathrm{ord}}\nolimits_\phi(e) = 2 \mathop{\mathrm{ord}}\nolimits_\phi(v)$, which happens if and only if $v$ is a
trivalent vertex, with three distinct edges $e,e'$ and $e''$ exiting from it, such that
$e' = \phi^{\mathop{\mathrm{ord}}\nolimits_\phi(v)}(e)$ and $\mathop{\mathrm{ord}}\nolimits_\phi(e'') = \mathop{\mathrm{ord}}\nolimits_\phi(v)$;
\@ifnextchar[{\item@}{\item@@}[3)] $\mathop{\mathrm{ord}}\nolimits_\phi(e) = 3
\mathop{\mathrm{ord}}\nolimits_\phi(v)$, which happens if and only if $v$ is a trivalent vertex, with three distinct
edges $e,e'$ and $e''$ exiting from it, such that $e' = \phi^{\mathop{\mathrm{ord}}\nolimits_\phi(v)}(e)$ and $e'' =
\phi^{2 \mathop{\mathrm{ord}}\nolimits_\phi(v)}(e)$.
\end{itemize}
Moreover, if $\Gamma$ is connected then the set of all orders of edges is $\{m, 2 m,
\dots, 2^k m\}$ for some $m \geq 1$ and $k \geq 0$ such that $\mathop{\mathrm{ord}}\nolimits(\phi) = 2^k m$, while
the set of all orders of vertices is $\{m, 2 m, \dots, 2^h m, 2^{j_1} m/3, \dots,
2^{j_\ell} m/3\}$ with $k-1 \leq h \leq k$, $0 \leq l \leq k$, $0 \leq j_1 < \dots <
j_\ell \leq k$, and $m$ multiple of $3$ if $\ell > 0$.
\end{lemma}
\begin{proof}
By definition of order for edges, we immediately have that $\mathop{\mathrm{ord}}\nolimits_\phi(e)$ is a multiple of
$\mathop{\mathrm{ord}}\nolimits_\phi(v)$. Moreover, since $\phi^{i \mathop{\mathrm{ord}}\nolimits_\phi(v)}(e)$ is an edge exiting from
$\phi^{i \mathop{\mathrm{ord}}\nolimits_\phi(v)}(v) = v$ for any $i \geq 1$, the only possible cases are
$\mathop{\mathrm{ord}}\nolimits_\phi(e) = \mathop{\mathrm{ord}}\nolimits_\phi(v), 2 \mathop{\mathrm{ord}}\nolimits_\phi(e), 3 \mathop{\mathrm{ord}}\nolimits_\phi(v)$ and these are verified
according to the conditions given in the statement. As a consequence, if the orders of two
adjacent edges are different, then one of them is the double of the other, being those
edges like $e$ and $e''$ in point~2. For $\Gamma$ connected, this implies that the set of
orders of the edges is $\{m, 2 m, \dots, 2^k m\}$ for some $m \geq 1$ and $k \geq 0$. In
order to get the set of orders of vertices, it suffices to observe that if $e$ and $e'$
are two edges exiting from $v$, such that $\mathop{\mathrm{ord}}\nolimits_\phi(e') = 2 \mathop{\mathrm{ord}}\nolimits_\phi(e)$, then
$\mathop{\mathrm{ord}}\nolimits_\phi(v) = \mathop{\mathrm{ord}}\nolimits_\phi(e)$, which gives the orders $m, 2 m, \dots, 2^h m$ with $k-1
\leq h \leq k$. On the other hand, tripodes of edges whose order has the form $2^j m/3$
may appear, like in point~3, for any $j = 0, \dots, k$.
\end{proof}
By a {\sl path} $\alpha$ between the (possibly coinciding) vertices $v$ and $w$ in a graph
$\Gamma$ we mean a (possibly non-simple) chain of edges having $v$ and $w$ as its ends. We
call the number $\mathop{\mathrm{len}}\nolimits(\alpha)$ of (non-necessarily distinct) edges in the chain the {\sl
length} of $\alpha$. Moreover, we denote by $\bar \alpha$ the reversed path.
\vskip\medskipamount
The next four lemmas concern minimal paths between vertices in a given $\phi$-orbit.
\begin{lemma}\label{paths-disjoint/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$. Given any (simple) path $\alpha \subset \Gamma$ of minimal
length among all paths joining any two different vertices of a given $\phi$-orbit, let $v$
and $v'$ the ends of $\alpha$ in that $\phi$-orbit. If $i \neq j$ and the four points
$\phi^i(v), \phi^i(v'), \phi^j(v)$ and $\phi^j(v')$ are all distinct, then the paths
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ are disjoint.
\end{lemma}
\begin{proof}
\begin{Figure}[b]{paths-disjoint-pf/fig}
\vskip-3pt
\fig{}{paths-disjoint-pf.eps}[-3pt]
{}{No common edge $e$ between $\phi^i(\alpha)$ and $\phi^j(\alpha)$ with four distinct
ends.}
\end{Figure}
Let $\alpha$ be a path as in the statement. By contradition, assume $\phi^i(\alpha)$ and
$\phi^j(\alpha)$ are not disjoint for some $i \neq j$ satisfying the required condition.
Then, they share at least a common internal edge $e$, and we have one of the two
situations depicted in Figure~\ref{paths-disjoint-pf/fig}, depending on the fact that $e$
is traversed in the same direction or not when going from $\phi^i(v)$ to $\phi^i(v')$
along $\phi^i(\alpha)$ and from $\phi^j(v)$ to $\phi^j(v')$ along $\phi^j(\alpha)$. Notice
that the paths $\beta_{i,1}$ and $\beta_{i,2}$ forming $\phi^i(\alpha) - \mathop{\mathrm{Int}}\nolimits(e)$ are not
necessarily disjoint from the paths $\beta_{j,1}$ and $\beta_{j,2}$ forming
$\phi^j(\alpha) - \mathop{\mathrm{Int}}\nolimits(e)$. In any case, we find a path shorter than $\alpha$ between
different vertices in the given $\phi$-orbit. Namely, such path is either $\beta_{i,1}
\bar \beta_{j,1}$ or $\beta_{i,2} \bar \beta_{j,2}$ in the left side case, while it is
either $\beta_{i,1} \bar \beta_{j,2}$ or $\beta_{i,2} \bar \beta_{j,1}$ in the right side
case.
\end{proof}
\begin{lemma}\label{paths-adjacent/thm}
Let $\phi: \Gamma \to \Gamma$, $\alpha \subset \Gamma$, $v$ and $v'$, be as in
Lemma~\ref{paths-disjoint/thm}. If $i \neq j$ with $\phi^j(v) = \phi^i(v')$ and
$\phi^j(v') \neq \phi^i(v)$, then the union $\phi^i(\alpha) \cup \phi^j(\alpha)$ has the
structure shown in Figure~\ref{paths-adjacent/fig}, where: 1)~$k \geq 1$; 2)~the paths
$\delta_{i,1}, \dots, \delta_{i,k} \subset \phi^i(\alpha)$ are disjoint from the paths
$\delta_{j,1}, \dots, \delta_{j,k} \subset \phi^j(\alpha)$; 3)~$\mathop{\mathrm{len}}\nolimits(\delta_{i,h}) =
\mathop{\mathrm{len}}\nolimits(\delta_{j,k-h+1}) \geq 1$ for every $h = 1, \dots, k$; 4)~$\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) =
\mathop{\mathrm{len}}\nolimits(\delta_{j,k}) \geq \mathop{\mathrm{len}}\nolimits(\alpha)/2$; 5) all the common paths shared by
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ have length $\geq 1$, except the terminal one ending
at $\phi^j(v) = \phi^i(v')$, which can have length zero.
\end{lemma}
\begin{Figure}[htb]{paths-adjacent/fig}
\vskip-6pt
\fig{}{paths-adjacent.eps}[-3pt]
{}{$\phi^i(\alpha) \cup\phi^j(\alpha)$
for $\phi^j(v) = \phi^i(v')$ and $\phi^j(v') \neq \phi^i(v)$.}
\end{Figure}
\begin{proof}
Look at Figure~\ref{paths-disjoint-pf/fig} assuming $\phi^j(v) = \phi^i(v')$ and
$\phi^j(v') \neq \phi^i(v)$. By a similar argument as in the proof of the previous lemma,
the minimality of $\alpha$ implies that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ cannot share
an edge $e$ like in the left side of the figure, while if $\phi^i(\alpha)$ and
$\phi^j(\alpha)$ share an edge $e$ like in the right side of the figure then
$\mathop{\mathrm{len}}\nolimits(\beta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\beta_{j,2}) \geq \mathop{\mathrm{len}}\nolimits(\alpha)/2$. This gives properties 1 and
4.
\begin{Figure}[htb]{paths-adjacent-pf/fig}
\fig{}{paths-adjacent-pf.eps}[-3pt]
{}{Common edges $e$ and $e'$ cannot occur in the same order along the paths
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ in Lemma~\ref{paths-adjacent/thm}.}
\end{Figure}
For property 2, it suffices to observe that, if $e$ and $e'$ are two common edges of
$\phi^i(\alpha)$ and $\phi^j(\alpha)$, then they occur in opposite orders along the two
paths starting from $\phi^i(v)$ and $\phi^i(v')$ respectively. Indeed, in the contrary
case, we would have $\phi^i(\alpha) = \beta_{i,1} e \beta_{i,2} e' \beta_{i,3}$ and
$\phi^j(\alpha) = \beta_{j,1} \bar e \beta_{j,2} \bar e' \beta_{j,3}$ like in
Figure~\ref{paths-adjacent-pf/fig}, where $\beta_i$'s and the $\beta_j$'s are possibly
empty and non-necessarily disjoint. Then, the minimality of $\phi^i(\alpha)$ would imply
$\mathop{\mathrm{len}}\nolimits(\beta_{i,2}) + 2 \leq \mathop{\mathrm{len}}\nolimits(\beta_{j,2})$, while the minimality of $\phi^j(\alpha)$
would imply $\mathop{\mathrm{len}}\nolimits(\beta_{j,2}) + 2 \leq \mathop{\mathrm{len}}\nolimits(\beta_{i,2})$, which would be absurd.
At this point, property~3 and 5 immediately follow from the minimality of $\phi^i(\alpha)$
and $\phi^j(\alpha)$ and from the trivalency of the vertices respectively.
\end{proof}
\begin{lemma}\label{paths-diagonal/thm}
Let $\phi: \Gamma \to \Gamma$, $\alpha \subset \Gamma$, $v$ and $v'$, be as in
Lemma~\ref{paths-disjoint/thm}. If $i \neq j$ with $\phi^j(v) = \phi^i(v')$ and
$\phi^j(v') = \phi^i(v)$, then the union $\phi^i(\alpha) \cup \phi^j(\alpha)$ has the
structure shown in Figure~\ref{paths-diagonal/fig}, where: 1)~$k \geq 0$; 2)~the paths
$\delta_{i,1}, \dots, \delta_{i,k} \subset \phi^i(\alpha)$ are disjoint from the paths
$\delta_{j,1}, \dots, \delta_{j,k} \subset \phi^j(\alpha)$; 3)~$\mathop{\mathrm{len}}\nolimits(\delta_{i,h}) =
\mathop{\mathrm{len}}\nolimits(\delta_{j,k-h+1}) \geq 1$ for every $h = 1, \dots, k$; 4) all the common paths shared
by $\phi^i(\alpha)$ and $\phi^j(\alpha)$ have length $\geq 1$, except both the terminal
ones that can have length zero.
Moreover, this may happen only when $\mathop{\mathrm{ord}}\nolimits(\phi)$ is even.
\end{lemma}
\begin{Figure}[htb]{paths-diagonal/fig}
\vskip-3pt
\fig{}{paths-diagonal.eps}[-3pt]
{}{$\phi^i(\alpha) \cup\phi^j(\alpha)$
for $\phi^i(v) = \phi^j(v')$ and $\phi^j(v) = \phi^i(v')$.}
\end{Figure}
\begin{proof}
The same arguments in the proof of Lemma~\ref{paths-adjacent/thm} still work here, to give
the first part of the statement. In fact, in that proof the assumption that $\phi^j(v')
\neq \phi^i(v)$ was only used to obtain $k \geq 1$ and property 4 (in the statement of
that lemma).
For the last sentence, we observe that the conditions $\phi^j(v) = \phi^i(v')$ and
$\phi^j(v') = \phi^i(v)$ imply that $\phi^{j-i}$ swaps $v$ and $v'$, which in turn implies
that $\mathop{\mathrm{ord}}\nolimits(\phi^{j-i})$, and hence $\mathop{\mathrm{ord}}\nolimits(\phi)$ as well, is even.
\end{proof}
\begin{lemma}\label{paths-doubled/thm}
Let $\phi: \Gamma \to \Gamma$, $\alpha \subset \Gamma$, $v$ and $v'$, be as in
Lemma~\ref{paths-disjoint/thm}. If $i \neq j$ with $\phi^i(v) = \phi^j(v)$ and $\phi^i(v')
= \phi^j(v')$, then the union $\phi^i(\alpha) \cup \phi^j(\alpha)$ has the structure shown
in Figure~\ref{paths-doubled/fig}, where: 1)~$k \geq 0$; 2)~the paths $\delta_{i,1},
\dots, \delta_{i,k} \subset \phi^i(\alpha)$ are disjoint from the paths $\delta_{j,1},
\dots, \delta_{j,k} \subset \phi^j(\alpha)$; 3)~$\mathop{\mathrm{len}}\nolimits(\delta_{i,h}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,h})
\geq 1$ for every $h = 1, \dots, k$; 4) all the common paths shared by $\phi^i(\alpha)$
and $\phi^j(\alpha)$ have length $\geq 1$, except both the terminal ones that can have
length zero.
Moreover, apart for the trivial case of $\phi^i(\alpha) = \phi^j(\alpha)$,
this may happen only when $\mathop{\mathrm{ord}}\nolimits(\phi)$ is either even or an odd multiple of 3, being in
the latter case $\phi^i(\alpha) = \delta_{i,1}$ and $\phi^j(\alpha) = \delta_{j,1}$.
\end{lemma}
\begin{Figure}[htb]{paths-doubled/fig}
\fig{}{paths-doubled.eps}[-3pt]
{}{$\phi^i(\alpha) \cup\phi^j(\alpha)$
for $\phi^i(v) = \phi^j(v)$ and $\phi^i(v') = \phi^j(v')$.}
\end{Figure}
\begin{proof}
The proof of the first part of the statement is completely analogous to that of the
previous lemma, except for the orderings of the $\delta_i$'s along $\phi^i(\alpha)$ and of
the $\delta_j$'s along $\phi^j(\alpha)$, which now coincide.
To prove the second part, assume that $\phi^i(\alpha) \neq \phi^j(\alpha)$ and
$\mathop{\mathrm{ord}}\nolimits(\phi)$ is odd. Then, we have $\phi^i(\alpha) = \delta_{i,1}$ and $\phi^j(\alpha) =
\delta_{j,1}$. Other\-wise, $\phi^{j-i}$ would fix some edge of $\phi^i(\alpha) \cap
\phi^j(\alpha)$ and $\mathop{\mathrm{ord}}\nolimits(\phi^{j-i})$ should be even by Lemma~\ref{orders/thm}, in
contrast with the oddness of $\mathop{\mathrm{ord}}\nolimits(\phi)$. For the same reason $\phi^{j-i}$ cannot swap
$\phi^i(\alpha)$ and $\phi^j(\alpha)$, then it cyclically permutes the three paths
$\phi^i(\alpha)$, $\phi^j(\alpha)$ and $\phi^{2j-i}(\alpha)$, which meet each other only
at their common ends $\phi^i(v)$ and $\phi^i(v')$. This implies that $\mathop{\mathrm{ord}}\nolimits(\phi^{j-i})$,
and hence $\mathop{\mathrm{ord}}\nolimits(\phi)$ as well, is a multiple of 3.
\end{proof}
\@startsection{section}{1}{\s@ctind{The case of order $p^m$ with $p$ prime greater than $3$%
\label{order-pm/sec}}
The aim of this section is to show that any automorphism $\phi \in \cal A_{g,b}$ with order
$\mathop{\mathrm{ord}}\nolimits(\phi) = p^m$ for a prime $p > 3$ is $F$-equivalent to the composition of two
automorphisms of order $2$.
Actually, we will prove this fact under the weaker assumption that $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is a
multiple of neither $2$ nor $3$. In fact, all the arguments are only based on the
following special properties, which hold for any automorphism $\phi: \Gamma \to \Gamma$
of a (possibly disconnected) uni/trivalent graph $\Gamma$, having such an order $n$:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] $\mathop{\mathrm{ord}}\nolimits_\phi(v) = \mathop{\mathrm{ord}}\nolimits_\phi(e) = n$ for any vertex $v$ and edge $e$ of $\Gamma$,
thanks to Lemma~\ref{orders/thm};
\@ifnextchar[{\item@}{\item@@}[2)] the situations of Lemmas~\ref{paths-diagonal/thm} and \ref{paths-doubled/thm}
cannot occur.
\end{itemize}
\vskip\medskipamount
\begin{lemma}\label{cycle-p/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is a multiple of neither $2$
nor $3$.\break If $\alpha \subset \Gamma$ is a (simple) path of minimal length among all
the paths joining any two distinct $\phi$-equi\-valent vertices, then $\cup_i\,
\phi^i(\alpha)$ is a disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup
\phi^{n/\ell - 1}(C)$ of $n/\ell$ simple cycles (possibly a single one, for $\ell = n$),
each given by a concatenation of $\ell$ images of $\alpha$ with $\ell$ a divisor of $n$
greater than $1$.\break In fact, there exists a positive integer $s = t\, n/\ell < n$
with $(t,\ell) = 1$, such that the cycle $C$ is given by $\alpha \,\phi^s(\alpha)
\cdots \phi^{(\ell - 1) s}(\alpha)$. Moreover, up to $F$-equivalence we can assume
$\alpha$ to consist of a single edge $a$ (see Figure \ref{cycle/fig}, where $a_i$ stands
for $\phi^{is}(a)$ and a similar notation is adopted for the $v_i$'s as well).
\end{lemma}
\begin{Figure}[htb]{cycle/fig}
\fig{}{cycle.eps}[-3pt]
{}{The form of the cycle $C$ in Lemma~\ref{cycle-p/thm}.}
\end{Figure}
\begin{proof}
Let $\alpha$ be a path as in the statement and let $v \neq v'$ be its ends. For any $i
\neq j \!\!\mod n$, let us consider the two paths $\phi^i(\alpha)$ and $\phi^j(\alpha)$,
and the four different situations described in Lemmas~\ref{paths-disjoint/thm} to
\ref{paths-doubled/thm}, which cover all the possibilities, thanks to the
$\phi$-equivalence of $v$ and $v'$.
The situations of Lemmas~\ref{paths-diagonal/thm} and \ref{paths-doubled/thm} cannot
occur, hence in particular $\phi^i(\alpha) \neq \phi^j(\alpha)$. Therefore,
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ can have non-empty intersection only as in
Lemma~\ref{paths-adjacent/thm}. In this situation, if $\phi^i(u)$ with $u \in \alpha$ is
the common end of $\delta_{i,1}$ and $\delta_{j,k}$, then $\phi^j(u) \in \delta_{j,k}$
since $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) \geq \mathop{\mathrm{len}}\nolimits(\alpha)/2$. Actually, $\phi^j(u)$
has to coincide with either $\phi^j(v')$ or $\phi^i(u)$, otherwise the global minimality
of $\alpha$ would be contradicted.
Since $\phi^j(u) = \phi^i(u)$ is impossible, being $\mathop{\mathrm{ord}}\nolimits_\phi(u) = n$, we are left with
the unique possibility $\phi^j(u) = \phi^j(v')$, that is $u = v'$. Then, we have
$\phi^i(\alpha) = \delta_{i,1}$ and $\phi^j(\alpha) = \delta_{j,k}$, which meet only at
their common end $\phi^j(v) = \phi^i(v')$. Moreover, such end cannot be shared by any
other $\phi^h(\alpha)$. Otherwise, either $\phi^i(\alpha)$ and $\phi^h(\alpha)$ or
$\phi^j(\alpha)$ and $\phi^h(\alpha)$ would be in the situation of
Lemma~\ref{paths-doubled/thm}.
Then, we can conclude in a straightforward way that $\cup_i\,\phi^i(\alpha)$ has the
stated form, with $s < n$ uniquely determined by $v' = \phi^s(v)$, $\ell$ the smallest
positive integer such that $\phi^{\ell s}(v) = v$, and $t = \ell\, s/n$ (this is an
integer since $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n$ and it is coprime with $\ell$ by the minimality of
$\ell\,$).
\begin{Figure}[htb]{cycle-pf/fig}
\fig{}{cycle-pf.eps}[-3pt]
{}{Reducing $\mathop{\mathrm{len}}\nolimits(\alpha)$ in Lemma \ref{cycle-p/thm}.}
\end{Figure}
At this point, we are left to prove that the path $\alpha$ can be replaced by a single
edge up to $F$-equivalence. Proceeding by induction, it suffices to show how to reduce
$\mathop{\mathrm{len}}\nolimits(\alpha)$ whenever this is greater than 1. Such reduction can be achieved by sliding
all the copies $f_i = \phi^{is}(f)$ of the unique edge $f$ attached at the first
intermediate vertex of $\alpha$ but not belonging to $\alpha$, so that their ends are
slided out of the paths $\alpha_i = \phi^{is}(\alpha)$, as indicated in Figure
\ref{cycle-pf/fig}. According to Remark \ref{slidings/rem}, those slidings correspond to a
$\phi$-invariant $F$-move, given by simultaneous elementary $F$-moves performed on the
first edges of all the $\alpha_i$'s, which are pairwise disjoint if $\mathop{\mathrm{len}}\nolimits(\alpha) \geq 2$.
\end{proof}
\begin{Figure}[b]{onion/fig}
\fig{}{onion.eps}[-3pt]
{}{The two possible forms of the cycle $C$ in Lemma~\ref{step-p/thm}.}
\end{Figure}
\begin{lemma}\label{step-p/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is a multiple of neither $2$
nor $3$.\break If $\alpha \subset \Gamma$ is a (simple) path of minimal length among all
paths joining any two distinct (hence disjoint, see proof) terminal edges in a given
$\phi$-orbit, then up to $F$-equivalence we can assume $\cup_i\,\phi^i(\alpha)$ to be
a disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$ of
$n/\ell$ cycles (possibly a single one, for $\ell = n$), with $\ell$ a divisor of $n$
greater than $1$ and $C = \alpha \,\phi^s(\alpha) \cdots \phi^{(\ell - 1) s}(\alpha)$ for
some positive integer $s = t\, n/\ell < n$ with $(t,\ell) = 1$, like in Lemma
\ref{cycle-p/thm}. Moreover, $\alpha$ can be assumed to
consist of either one edge $a$ or two edges $a$ and $b$ (see Figure \ref{onion/fig}, where
$a_i$ and $b_i$ stand for $\phi^{is}(a)$ and $\phi^{is}(b)$ respectively, and a similar
notation is adopted for the terminal edges $e_i$'s and the vertices $v_i$'s as well).
\end{lemma}
\begin{proof}
Let $\alpha$ be a path as in the statement, and $e \neq e'$ be the terminal edges it
joins. Then the ends of $\alpha$ coincide with the unique trivalent ends $v$ and $v'$
of $e$ and $e'$ respectively. Notice that $v \neq v'$, being $\mathop{\mathrm{ord}}\nolimits_\phi(v) = \mathop{\mathrm{ord}}\nolimits_\phi(e)
= n$ by Lemma \ref{orders/thm}.
For any $i \neq j \!\!\mod n$, by arguing as in the proof of Lemma~\ref{cycle-p/thm}, we
can prove that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can meet only as in
Lemma~\ref{paths-adjacent/thm}, and that their common end $\phi^j(v) = \phi^i(v')$ cannot
be shared by any other $\phi^h(\alpha)$. The same Lemma \ref{paths-adjacent/thm} also
tells us that $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) \geq \mathop{\mathrm{len}}\nolimits(\alpha)/2$ (cf.
Figure~\ref{paths-adjacent/fig}). But here the equality cannot occur, otherwise
$\phi^{j-i}$ would fix the common end of $\delta_{i,1}$ and $\delta_{j,k}$, and it would
cyclically permute the three edges exiting from it, in contrast with the assumption that
$n$ is not a multiple of 3.
Then, we can conclude straightforwardly that $\cup_i\,\phi^i(\alpha)$ consists of the
disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$, with $s < n$
uniquely determined by $\phi^s(v) = v'$, $\ell$ the smallest positive integer such that
$\phi^{\ell s}(v) = v$, $t = s\, \ell/n$ (cf. proof of Lemma \ref{cycle-p/thm}), and $C =
\alpha \,\phi^s(\alpha) \cdots \phi^{(\ell - 1) s}(\alpha)$ having the form depicted in
Figure \ref{onion-pf1/fig}. Here, and in the next figure as well, $\alpha_i$ stands for
the path $\phi^{is}(\alpha)$, which is the part of $C$ in the corresponding circular
sector, and each arc represents a path of edges.
\begin{Figure}[htb]{onion-pf1/fig}
\vskip6pt
\fig{}{onion-pf1.eps}
{}{The starting configuration $C$ in the proof of Lemma \ref{step-p/thm}.}
\end{Figure}
In the very special case where the paths $\alpha_i$ consist of single edges, they can only
share their common ends. Hence, the configuration $C$ in Figure \ref{onion-pf1/fig}
coincides with that in the left side of Figure \ref{onion/fig}, and we are done.
Otherwise, by a sequence of $\phi$-invariant slidings (cf. Remark \ref{slidings/rem}) on
the edges attached to $C$ and to its copies, as indicated in Figure \ref{onion-pf2/fig}
for three of such edges, we can reduce all the arcs in Figure \ref{onion-pf1/fig} to
single edges, possibly except the ones forming the big circle. In the same way, when the
first edge of $\alpha_i$ coincides with the last edge of $\alpha_{i-1 \!\!\mod \ell}$, we
also fuse it with the edge $e_i = \phi^{is}(e)$.
\begin{Figure}[htb]{onion-pf2/fig}
\fig{}{onion-pf2.eps}[-3pt]
{}{Simplifying the starting configuration.}
\end{Figure}
Then, we can eliminate all the resulting bigons, by a $\phi$-invariant $F$-move, which
atcs on $C$ by simultaneous elementary $F$-moves on the edges between any two consecutive
of them and between the most external ones and the big circle, as shown in Figure
\ref{onion-pf3/fig}.
\begin{Figure}[t]{onion-pf3/fig}
\fig{}{onion-pf3.eps}[-3pt]
{}{Eliminating the bigons by simultaneous edge $F$-moves.}
\end{Figure}
\begin{Figure}[htb]{onion-pf4/fig}
\fig{}{onion-pf4.eps}
{}{Reducing $\mathop{\mathrm{len}}\nolimits(\alpha)$ in Lemma \ref{step-p/thm}.}
\end{Figure}
Finally, to get the configuration in the right side of Figure \ref{onion/fig}, it remains
to reduce $\alpha$ to a chain of only two edges $a$ and $b$. This can be done proceeding
by induction, like in the last part of the proof of Lemma \ref{cycle-p/thm}. In Figure
\ref{onion-pf4/fig} we see how to reduce $\mathop{\mathrm{len}}\nolimits(\alpha)$ whenever this is greater than 2.
Here, $a_i = \phi^{is}(a)$ and $f_i = \phi^{is}(f)$ are respectively the images in
$\alpha_i = \phi^{is}(\alpha)$ of the first edge $a$ of $\alpha$ and of the edge $f$
attached at the second intermediate vertex of $\alpha$ but not belonging to $\alpha$.
The desired reduction is achieved through an invariant $F$-move, which simultaneously
slides all the $f_i$, in such a way that their ends pass from the $\alpha_i$'s to the
first edges attached to them. Notice that, contrary to what we did in the proof of Lemma
\ref{cycle-p/thm}, we cannot further reduce $\alpha$ to a single edge by a slide letting
even the last intermediate point of $\alpha$ pass to $e$.
\end{proof}
\begin{proposition}\label{onion-p/thm}
Let $\phi \in \cal A_{g,b}$ be a non-trivial automorphism, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi) > 1$ is
a multiple of neither $2$ nor $3$. Then, $\phi$ is $F$-equivalent to the composition $\tau
\circ \sigma$ of two automorphisms $\sigma, \tau \in \cal A_{g,b}$ such that $\mathop{\mathrm{ord}}\nolimits(\sigma) =
\mathop{\mathrm{ord}}\nolimits(\tau) = 2$. Moreover, both $\sigma$ and $\tau$ can be assumed to fix an edge and
reverse an invariant edge.\hfill
\end{proposition}
\begin{proof}
Let $\phi: \Gamma \to \Gamma$ any automorphism in $\cal A_{n,g}$ as in the statement. We will
construct by recursion a sequence $\phi_i: \Gamma_i \to \Gamma_i$ of automorphisms in
$\cal A_{g,b}$ and a sequence of subgraphs $\Lambda_i \subset \Gamma_i$ with $i = 0, \dots, r$
and $r \geq 2$, having the following properties:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] $\phi_0 = \phi$ (hence $\Gamma_0 = \Gamma$), and $\phi_i$ is $F$-equivalent to
$\phi_{i-1}$ for every $1 \leq i \leq r$;
\@ifnextchar[{\item@}{\item@@}[2)] $\emptyset = \Lambda_0 \varsubsetneq \Lambda_1 \varsubsetneq \dots \varsubsetneq
\Lambda_r = \Gamma_r$, and $\Delta_i = \mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_i)$ is a uni/trivalent graph
whose intersection with $\Lambda_i$ is a single $\phi_i$-orbit $U_i \subset \Delta_i$ of
its free ends for every $1 \leq i \leq r-1$;
\@ifnextchar[{\item@}{\item@@}[3)] $\Lambda_i$ is $\phi_i$-invariant and $\phi_{i|\Lambda_{i-1}} =
\phi_{i-1|\Lambda_{i-1}}$ for every $1 \leq i \leq r$;
\@ifnextchar[{\item@}{\item@@}[4)] $\phi_{i|\Lambda_i}: \Lambda_i \to \Lambda_i$ is the composition $\tau_i \circ
\sigma_i$ of two automorphisms $\sigma_i: \Lambda_i \to \Lambda_i$ and $\tau_i: \Lambda_i
\to \Lambda_i$ such that $\mathop{\mathrm{ord}}\nolimits(\sigma_i) = \mathop{\mathrm{ord}}\nolimits(\tau_i) = 2$, for every $1 \leq i \leq r$;
moreover, $\sigma_i$ fixes exactly one vertex $u_i$ in $U_i$ (hence $\tau_i$ fixes
$\phi_i^{(n+1)/2}(u_i) \in U_i$) for every $1 \leq i \leq r-1$, and both $\sigma_i$ and
$\tau_i$ fix an edge and reverse an invariant edge for every $2 \leq i \leq r$.
\end{itemize}
This will prove the proposition, since $\sigma = \sigma_r$ and $\tau = \tau_r$ satisfy
all the required conditions, being $\tau \circ \sigma = \tau_r
\circ \sigma_r = \phi_{r|\Lambda_r} = \phi_r$ $F$-equivalent to $\phi$ in $\cal A_{g,b}$.
To start the construction, we apply Lemma \ref{cycle-p/thm} to the automorphism $\phi$
and a minimal path $\alpha \subset \Gamma$ as in that statement (this exists since
$\Gamma$ is connected but not reduced to a single vertex and $\phi$ is non-trivial). As a
result, we get an automorphism $\phi_1: \Gamma_1 \to \Gamma_1$ in $\cal A_{g,b}$, which is
$F$-equivalent to $\phi_0 = \phi$ and such that the minimal path corresponding to $\alpha$
in $\Gamma_1$ consists of only one (internal) edge $a \subset \Gamma_1$. Then, we put
$\Lambda_1 = \cup_j\, \phi_1^j(a) \subset \Gamma_1$, which is $\phi_1$-invariant by
definition. Lemma \ref{cycle-p/thm} tells us that there exist a divisor $\ell > 1$ of $n$
and an integer $0 < t < l$ with $(t,\ell) = 1$, such that $\Lambda_1$ consists of the
disjoint union $C \sqcup \phi_1(C) \sqcup \dots \sqcup \phi_1^{n/\ell - 1}(C)$ of $n/\ell$
simple cycles (possibly a single one, for $\ell = n$), where $C = a \,\phi_1^s(a) \cdots
\phi_1^{(\ell - 1) s}(a)$\break with $s = t\, n/\ell$.
Let $v$ be the first end of the edge $a$, so that $\phi_1^s(v)$ is the second one. Then,
the set of vertices of $\Lambda_1$ is the $\phi_1$-orbit $V = \{v, \phi_1(v), \dots,
\phi_1^{n-1}(v)\}$, and any automorphism on $\Lambda_1$ is uniquely determined by its
action on $V$, since $n$ is odd and hence $\ell > 2$. Therefore, we can first define the
two involutions $\sigma_1$ and $\tau_1$ on $V$ in such a way that $\tau_1 \circ \sigma_1$
coincides with the restriction of $\phi_1$ to $V$, then check that they preserve
adjacency of vertices in $\Lambda_1$ and hence give well-defined automorphisms of
$\Lambda_1$.
We define $\sigma_1$ and $\tau_1$ on $V$ by putting $\sigma_1(\phi_1^j(v)) =
\phi_1^{-j}(v)$ and $\tau_1(\phi_1^j(v)) = \phi_1^{1-j}(v)$, for $j = 0, \dots, n-1$.
Clearly, these are involutions and their composition $\tau_1 \circ \sigma_1$ acts on $V$
as $\phi_1$. Concerning the preservation of adjacency, it suffices to observe that two
elements $\phi_1^j(v)$ and $\phi_1^k(v)$ of $V$ are adjacent vertices in $\Lambda_1$ if
and only if $|j - k| = s \!\!\mod n$, and both $\sigma_1$ and $\tau_1$ preserve this
condition.
Finally, we note that $\Delta_1 = \mathop{\mathrm{Cl}}\nolimits(\Gamma_1 - \Lambda_1)$ is a uni/trivalent graph and
$\Delta_1 \cap \Lambda_1$ is given by the $\phi_1$-orbit $U_1 = V$ of free ends of
$\Delta_1$, as required in point~2 above. Moreover, $\sigma_1$ fixes only the vertex $u_1
= v$ of $U_1$ and reverses the edge $\phi_1^{(n+1)/2}(a)$, while $\tau_1$ reverses the
edge $a$, as required in point~4 above.
Now, to realize the recursive step of the construction, assume we are given $\phi_{i-1}:
\Gamma_{i-1} \to \Gamma_{i-1}$, $u_{i-1} \in U_{i-1} \subset \Lambda_{i-1} \varsubsetneq
\Gamma_{i-1}$ and $\sigma_{i-1},\tau_{i-1}: \Lambda_{i-1} \to \Lambda_{i-1}$ satisfying
all the properties stated in the above points~1 to 4 for $i-1 < r$.\break Lemmas
\ref{ordinv/thm} and \ref{orders/thm} ensure that $\mathop{\mathrm{ord}}\nolimits(\phi_{i-1}) = n$ and that the
cardinality of the $\phi_{i-1}$-orbit $U_{i-1}$ is equal to $n$ as well. Hence, also the
restriction of $\phi_{i-1}$ to the uni/tri\-valent graph $\Delta_{i-1} = \mathop{\mathrm{Cl}}\nolimits(\Gamma_{i-1}
- \Lambda_{i-1})$ is an automorphism of the same order $n$. Furthermore, the terminal
edges of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$ are all distinct. In fact, if two
of them would coincide, then their $\phi$-order would be even, which contradicts the
hypothesis that $n$ is odd.
If $\Delta_{i-1}$ does not contain any path joining different terminal edges ending at
vertices in $U_{i-1}$, then it consists of $n$ components each containing a single vertex
in $U_{i-1}$. Denoting by $C$ the component of $\Delta_{i-1}$ containing that vertex
$u_{i-1}$, we have the component decomposition $\Delta_{i-1} = C \sqcup \phi_{i-1}(C)
\sqcup \dots \sqcup \phi_{i-1}^{n-1}(C)$, with $\phi_{i-1}$ cyclically permuting the
components. In this case, we put $\phi_i = \phi_{i-1}$ and $\Lambda_i = \Gamma_i =
\Gamma_{i-1}$. Moverover, we define $\sigma_i$ and $\tau_i$ as the unique automorphisms of
$\Gamma_i$ extending $\sigma_{i-1}$ and $\tau_{i-1}$, in such a way that they act on each
$\phi_{i-1}^j(C)$ with $j = 0, \dots, n-1$, as $\phi_{i-1}^{-2j}$ and $\phi_{i-1}^{1-2j}$
respectively. A straightforward verification shows that such $\phi_i$, $\Lambda_i$,
$\sigma_i$ and $\tau_i$ satisfy all the properties stated in the above points~1 to 4 for
the case when $i = r$, which means that this step terminates the recursion.
Otherwise, if a path $\alpha \subset \Delta_{i-1}$ exists joining different terminal edges
of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$, then we can choose it to be minimal
(hence simple) and apply Lemma \ref{step-p/thm} to the restriction of $\phi_{i-1}$ to
$\Delta_{i-1}$ and to such a minimal $\alpha$, in order to get the structure described in
that statement for $\cup_j\, \phi_{i-1}^j(\alpha) \subset \Delta_{i-1}$ up to
$F$-equivalence. Such an $F$-equivalence only involves internal edges of $\Delta_{i-1}$,
hence it does not change the set of free ends $U_{i-1} \subset \Delta_{i-1}$ and the
restriction of $\phi_{i-1|\Delta_{i-1}}$. Therefore, it can be extended to an
$F$-equivalence of the whole $\phi_{i-1}$ on $\Gamma_{i-1}$, which leaves $\Lambda_{i-1}$
and the restriction $\phi_{i-1|\Lambda_{i-1}}$ unchanged.
As a result we get a new uni/trivalent graph $\Gamma_i$, such that $\Lambda_{i-1} \subset
\Gamma_i$ and a new automorphism $\phi_i: \Gamma_i \to \Gamma_i$, which is $F$-equivalent
to $\phi_{i-1}$ and coincides with $\phi_{i-1}$ on $\Lambda_{i-1}$. We also get a new
minimal path $\alpha \subset \mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$, which join different
$\phi_i$-equivalent terminal edges of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ ending at vertices
in $U_{i-1}$, such that $\cup_j\, \phi_i^j(\alpha)$ itself (no more up to $F$-equivalence)
has the structure described in Lem\-ma~\ref{step-p/thm}. Let $e \subset \Gamma_i$ be the
terminal edge of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ ending at $u_{i-1}$, and $v \in
\Gamma_i$ be the other end of $e$, which is a trivalent vertex of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i -
\Lambda_{i-1})$. Up to $\phi_i$-equiv\-alence, we can assume that $\alpha \subset
\Gamma_i$ starts at $v$ and ends at $\phi_i^s(v)$ with $0 < s < n$. According to Lemma
\ref{step-p/thm}, we can also assume that $\alpha$ consists of either one edge $a$ or two
edges $a$ and $b$ sharing the vertex $u$, while $\cup_j\, \phi_i^j(\alpha)$ consists of
the disjoint union $C \sqcup \phi_i(C) \sqcup \dots \sqcup \phi_i^{n/\ell - 1}(C)$ of
$n/\ell$ cycles (possibly a single one, for $\ell = n$), where $C = \alpha
\,\phi_i^s(\alpha) \cdots \phi_i^{(\ell - 1) s}(\alpha)$ for some $\ell > 1$ such that $s
\ell = t n$ with $(t,\ell) = 1$.
We put $\Lambda_i = \Lambda_{i-1} \cup_j\, \phi_i^j(e) \cup_j\, \phi_i^j(\alpha)$ and
define $\sigma_i$ and $\tau_i$ to be the unique automorphisms of $\Lambda_i$ extending
$\sigma_{i-1}$ and $\tau_{i-1}$ respectively. To see that such extensions exist, we first
define $\sigma_i(\phi_i^j)(v) = \phi_i^{-j}(v)$ and $\tau_i(\phi_i^j)(v) =
\phi_i^{1-j}(v)$, and also $\sigma_i(\phi_i^j)(u) = \phi_i^{-j}(u)$ and
$\tau_i(\phi_i^j)(u) = \phi_i^{1-j}(u)$ in the case when $\alpha = ab$, for every $j =
0, \dots, n-1$. Then, essentially by the same argument used above for $i = 1$, we check
that these definitions, together with $\sigma_i(\phi_i^j)(u_{i-1}) = \phi_i^{-j}(u_{i-1})$
and $\tau_i(\phi_i^j)(u_{i-1}) = \phi_i^{1-j}(u_{i-1})$ for every $i=0, \dots, n$,
preserve adjacency of vertices in $\Lambda_i$, hence they determine automorphisms
$\sigma_i$ and $\tau_i$ of the graph $\Lambda_i$ extending $\sigma_{i-1}$ and $\tau_{i-1}$
respectively. Notice that $\tau_i \circ \sigma_i$ trivially coincides with
$\phi_{i|\Lambda_i}$. Moreover, $\sigma_i$ fixes the edge $e$ and reverses the edge
$\phi_i^{(n+1)/2}(a)$, while $\tau_i$ fixes the edge $\phi_i^{(n+1)/2}(e)$ and reverses
the edge $a$. Finally, $u_i = \phi_i^{(n+1)/2}(u)$ is the unique vertex in its
$\phi_i$-orbit $U_i = \{\phi_i^j(u)\,|\,j = 0, \dots, n\}$ fixed by $\sigma_i$.
At this point, if $\alpha$ consists of the only edge $a$, hence $C$ has the structure
depicted in the left side of Figure \ref{onion/fig}, then $\Lambda_i = \Gamma_i$ by the
connectedness of $\Gamma_i$ and the recursion terminates with $r = i$. Otherwise, if
$\alpha$ consists of the two edges $a$ and $b$, hence $C$ has the structure depicted in
the right side of Figure \ref{onion/fig}, then we put $\Delta_i = \mathop{\mathrm{Cl}}\nolimits(\Gamma_i -
\Lambda_i)$ and observe that this is a uni/trivalent graph such that $\Delta_i \cap
\Lambda_i = U_i$. This conclude the recursive step.
\end{proof}
\@startsection{section}{1}{\s@ctind{The case of order $3^m$%
\label{order-3m/sec}}
In this section, we want to prove that any automorphism $\phi \in \cal A_{g,b}$ with order
$\mathop{\mathrm{ord}}\nolimits(\phi) = 3^m$ is $F$-equivalent to the composition of two automorphisms of order $2$.
Like in the prevous section, a weaker assumption will suffice to get the same conclusion.
Namely, we will assume that $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is an odd multiple of $3$. The only
properties we will need, for any automorphism $\phi: \Gamma \to \Gamma$ of a (possibly
discon\-nected) uni/trivalent graph $\Gamma$ with such an order $n$, are the following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] $\mathop{\mathrm{ord}}\nolimits_\phi(e) = n$ for any edge $e$ of $\Gamma$, while $\mathop{\mathrm{ord}}\nolimits_\phi(v)$ is either
$n$ or $n/3$ for any vertex $v$ of $\Gamma$, thanks to Lemma~\ref{orders/thm};
\@ifnextchar[{\item@}{\item@@}[2)] the situation of Lemmas~\ref{paths-diagonal/thm} cannot occur, while that of
Lemma \ref{paths-doubled/thm} can only occur with the two paths $\phi^i(\alpha)$ and
$\phi^j(\alpha)$ sharing no edge.
\end{itemize}
The whole argument is essentially the same as in the previous section, except for some
more cases occurring here. Hence, in both statements and proofs we will just concern with
those extra cases, while referring for the others to the analogous statements and proofs
in the previous section.
\begin{lemma}\label{cycle-3/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is an odd multiple of $3$. If
$\alpha \subset \Gamma$ is a (simple) path of minimal length among all paths joining any
two distinct $\phi$-equivalent vertices, then $\cup_i\, \phi^i(\alpha)$ is one
of the following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] a disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$
of $n/\ell$ simple cycles (possibly a single one, for $\ell = n$), each given by a
concatenation of images of $\alpha$, having the same structure as described in Lemma
\ref{cycle-p/thm}; in this case, up to $F$-equivalence we can assume $\alpha$ to
consist of a single edge $a$ (see Figure \ref{cycle/fig}, where $a_i$ stands for
$\phi^{is}(a)$ and a similar notation is adopted for the vertices $v_i$ as well);
\@ifnextchar[{\item@}{\item@@}[2)] a disjoint union $T \sqcup \phi(T) \sqcup \dots \sqcup \phi^{n/3\,-\,1}(T)$ of
$n/3$ tripods (possibly a single one, for $n=3$), where $T = \alpha \cup
\phi^{n/3}(\alpha) \cup \phi^{2n/3}(\alpha)$ and $\alpha$ is the concatenation of two
edges $a$ and $b$, such that $b = \phi^{n/3}(\bar a)$, $\phi^{n/3}(b) = \phi^{2n/3}(\bar
a)$ and $\phi^{2n/3}(b) = \bar a$ (see Figure~\ref{tripode1/fig}, where $a_i$ stands for
$\phi^{i \mkern1mu n/3}(a)$ and a similar notation is adopted for the vertices $v_i$,
which can be either all univalent as on the left side or all trivalent as on the right
side).
\end{itemize}
\end{lemma}
\begin{Figure}[htb]{tripode1/fig}
\vskip-21pt
\fig{}{tripode1.eps}[-3pt]
{}{The form of the tripode $T$ in Lemma \ref{cycle-3/thm}.}
\end{Figure}
\begin{proof}
Let $\alpha \subset \Gamma$ be an arbitrary path of minimal length among all paths joining
any two distinct $\phi$-equivalent vertices and let $v \neq v'$ be its ends. For any $i
\neq j \!\!\mod n$, let us consider the two paths $\phi^i(\alpha)$ and $\phi^j(\alpha)$,
and the four different situations described in Lemmas~\ref{paths-disjoint/thm} to
\ref{paths-doubled/thm}, which cover all the possibilities, thanks to the
$\phi$-equivalence of $v$ and $v'$.
The situation of Lemma~\ref{paths-diagonal/thm} cannot occur, while that of Lemma
\ref{paths-doubled/thm} could possibly occur only with $\phi^i(\alpha)$ and
$\phi^j(\alpha)$ sharing no edge, but the argument below shows this is actually
impossible.
Assume by contradiction that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ are as in Lemma
\ref{paths-doubled/thm}, hence they meet at $\phi^i(v) = \phi^j(v)$ and $\phi^i(v') =
\phi^j(v')$. Let us denote by $a$ the first edge of $\alpha$ starting from $v$ and by $u$
the other end of $a$. Then, the vertices $\phi^i(u)$ and $\phi^j(u)$ are distinct since
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ do not share any edge. Being $n$ an odd multiple of
3, thanks to Lemma~\ref{orders/thm} we have $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n/3$ and $\mathop{\mathrm{ord}}\nolimits_\phi(u)= n$.
As a consequence, we get $\mathop{\mathrm{len}}\nolimits(\alpha) = 2$. In fact, the concatenation $\phi^i(\bar a)\,
\phi^j(a)$ is simple path of length $2$ between the two distinct $\phi$-equivalent
vertices $\phi^i(u)$ and $\phi^j(u)$. Hence, by the global minimality of $\alpha$ we have
$\mathop{\mathrm{len}}\nolimits(\alpha) \leq 2$. On the other hand, $\alpha$ cannot be reduced to the single edge
$a$, otherwise $v$ and $u$ should be $\phi$-equivalent, in contrast with $\mathop{\mathrm{ord}}\nolimits_\phi(u)
\neq \mathop{\mathrm{ord}}\nolimits_\phi(v)$.
Now, let $s$ be an integer such that $v' = \phi^s(v)$. Then, $\phi^s(u)$ coincides with
one of $u$, $\phi^{n/3}(u)$ and $\phi^{2n/3}(u)$, since these are the three vertices
adjacent to $v'$. In any case, we can conclude that $s$ is a multiple of $n/3$, which is
in contrast with $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n/3$ and $v' \neq v$. This proves that the situation of
Lemma \ref{paths-doubled/thm} cannot occur.
At this point, we have that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can only meet as in
Lemma~\ref{paths-adjacent/thm}. Let $\phi^i(u)$ with $u \in \alpha$ be the common end
of $\delta_{i,1}$ and $\delta_{j,k}$. As in the proof of Lemma~\ref{cycle-p/thm}, we can
see that either $\phi^j(u) = \phi^j(v')$ or $\phi^j(u) = \phi^i(u)$, due to the global
minimality of $\alpha$. But differently from that proof, here the latter equality can
actually hold when $u$ has order $n/3$ and $j = i \!\!\mod n/3$.
If $\phi^j(u) = \phi^i(u)$, then the global minimality of $\alpha$ implies that both
$\delta_{i,1}$ and $\delta_{j,k}$ have length 1. Therefore, $\alpha$ is the concatenation
of two edges $a$ and $b$ sharing the vertex $u$, and hence $T = \alpha \cup
\phi^{n/3}(\alpha) \cup \phi^{2n/3}(\alpha)$ has the form described in point~2 of
the statement. Of course, this means that $\cup_i\,\phi^i(\alpha) = T \cup \phi(T) \cup
\dots \cup \phi^{n/3\,-\,1}(T)$. Moreover, any two different copies of $T$ in this union
are disjoint, otherwise they should share a free end, and there would be two paths
$\phi$-equivalent to $\alpha$ exiting from that common vertex like in
Lemma~\ref{paths-doubled/thm}.
Notice that the last argument shows that, if $\phi^j(u) = \phi^i(u)$ for some $i$ and $j$,
then the same holds for any $i$ and $j$ such that $\phi^i(\alpha)$ and $\phi^j(\alpha)$
are not disjoint.
So, we are left with the case when $\phi^j(u) = \phi^j(v')$ for any two non-disjoint paths
$\phi^i(\alpha)$ and $\phi^j(\alpha)$. In this case, the same argument exploited in the
proof of Lemma~\ref{cycle-p/thm}, allows us to conclude that the situation is the one
described in point~1 of the statement.
\end{proof}
\begin{lemma}\label{step-3/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi)$ is an odd multiple of $3$. If
$\alpha \subset \Gamma$\break is a (simple) path, possibly degenerate to a single vertex,
of minimal length among all paths joining any two distinct terminal edges in a given
$\phi$-orbit, then up to $F$-equivalence we can assume $\cup_i\,\phi^i(\alpha)$ to
be one of the following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] a disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$
of $n/\ell$ simple cycles (possibly a single one, for $\ell = n$), each given by a
concatenation of images of $\alpha$, having the same structure as described in
Lemma~\ref{step-p/thm}, with $\alpha$ consisting of either one edge $a$ or two edges $a$
and $b$ (see Figure~\ref{onion/fig}, where $a_i$ and $b_i$ stand for $\phi^{is}(a)$ and
$\phi^{is}(b)$ respectively, and a similar notation is adopted for the terminal edges
$e_i$'s and the vertices $v_i$'s as well);
\@ifnextchar[{\item@}{\item@@}[2)] a disjoint union $T \sqcup \phi(T) \sqcup \dots \sqcup \phi^{n/3\,-\,1}(T)$ of
$n/3$ tripods (possibly a single one, for $n=3$), where $T = \alpha \cup
\phi^{n/3}(\alpha) \cup \phi^{2n/3}(\alpha)$, with $\alpha$ being the concatenation of two
edges $a$ and $b$, such that $b = \phi^{n/3}(\bar a)$, $\phi^{n/3}(b) = \phi^{2n/3}(\bar
a)$ and $\phi^{2n/3}(b) = \bar a$ (see left side of Figure~\ref{tripode2/fig}, where $a_i$
and $e_i$ stand for $\phi^{i \mkern1mu n/3}(a)$ and $\phi^{i \mkern1mu n/3}(e)$
respec\-tively);
\@ifnextchar[{\item@}{\item@@}[3)] a set of $n/3$ trivalent vertices, when $\alpha$ reduces to a single vertex,
that is the terminal edges it joins share a common trivalent end; in this case the whole
graph $\Gamma$ consists of a disjoint union $\hat T \sqcup \phi(\hat T) \sqcup \dots
\sqcup \phi^{n/3\,-\,1}(\hat T)$ of $n/3$ tripods, where $\hat T = e \cup \phi^{n/3}(e)
\cup \phi^{2n/3}(e)$ with $\alpha$ reduced to the trivalent vertex of $\hat T$\break (see
right side of Figure~\ref{tripode2/fig}, where $e_i$ stands for the terminal edge $\phi^{i
\mkern1mu n/3}(e)$).
\end{itemize}
\end{lemma}
\begin{Figure}[htb]{tripode2/fig}
\vskip-18pt
\fig{}{tripode2.eps}
{}{The form of the tripodes $T$ (left) and $\hat T$ (right) in Lemma \ref{step-3/thm}.}
\end{Figure}
\begin{proof}
Let $\alpha$ be a path as in the statement and let $e \neq e'$ be the terminal edges it
joins. Then the ends of $\alpha$ coincide with the unique trivalent ends $v$ and $v'$
of $e$ and $e'$ respectively.
If $v = v'$, that is $\mathop{\mathrm{len}}\nolimits(\alpha) = 0$, then $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n/3$ according to
Lemma~\ref{orders/thm} and $e'$ is either $\phi^{n/3}(e)$ or $\phi^{2n/3}(e)$. In any
case, $\phi^{n/3}$ cyclically permutes the three edges of the tripode $\hat T = e \cup
\phi^{n/3}(e) \cup \phi^{2n/3}(e)$, and we have the situation described in point~3 of the
statement.
If $v \neq v'$, that is $\mathop{\mathrm{len}}\nolimits(\alpha) \geq 1$, we consider any two distinct paths
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ with $i \neq j \!\!\mod n$ and look once again at
the four possible situations described in Lemmas~\ref{paths-disjoint/thm} to
\ref{paths-doubled/thm}.
As in the proof of Lemma~\ref{cycle-3/thm}, the situations of
Lemmas~\ref{paths-diagonal/thm} and \ref{paths-doubled/thm} cannot occur. But here the
argument to exclude the latter is different. Namely, if $\phi^i(\alpha)$ and
$\phi^j(\alpha)$ were as in Lemma~\ref{paths-doubled/thm}, then $\phi^{j-i}$ should swap
the first edges of them starting from $\phi^i(v) = \phi^j(v)$, being $\phi^i(e) =
\phi^j(e)$ the third edge at that vertex, and this would be in contrast with the oddness
of $n$.
Therefore, $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can meet only as in
Lemma~\ref{paths-adjacent/thm}, and their common end $\phi^j(v) = \phi^i(v')$ cannot be
shared by any other $\phi^h(\alpha)$ (cf. proof of Lemma~\ref{cycle-p/thm}). Then,
consider the two subpaths $\delta_{i,1} \subset \phi^i(\alpha)$ and $\delta_{j,k} \subset
\phi^j(\alpha)$ in Figure~\ref{paths-adjacent/fig}.
\begin{Figure}[b]{tripode3/fig}
\vskip3pt
\fig{}{tripode3.eps}
{}{The starting configuration $T$ in the proof of Lemma \ref{cycle-3/thm}.}
\end{Figure}
If $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) = \mathop{\mathrm{len}}\nolimits(\alpha)/2$, then for the middle vertex
$u$ of $\alpha$ we have that $\phi^i(u) = \phi^j(u)$ is the common end of $\delta_{i,1}$
and $\delta_{j,k}$. By the same argument as above, this implies that $\mathop{\mathrm{ord}}\nolimits_\phi(u) = n/3$
and $\phi^{n/3}$ cyclically permutes the three paths $\alpha$, $\phi^{n/3}(\alpha)$ and
$\phi^{2n/3}(\alpha)$. Moreover, such three paths are disjoint from any other path
$\phi^h(\alpha)$, hence we can conlcude that $\cup_i\,\phi^i(\alpha)$ is a disjoint
union $T \sqcup \phi(T) \sqcup \dots \sqcup \phi^{n/3\,-\,1}(T)$, with $T$ as in the left
side of Figure~\ref{tripode3/fig}. Here, $\alpha_i$ stands for the path $\phi^{i \mkern1mu
n/3}(\alpha)$, which is the part of $T$ in the corresponding circular sector, and any arc
except the terminal ones represent a path of edges. Analogously, $v_i$ and $e_i$ stand for
$\phi^{i \mkern1mu n/3}(v)$ and $\phi^{i \mkern1mu n/3}(e)$ respectively.
We can reduce any arc in this starting configuration $T$ to a single edge,
by performing $\phi$-invariant slidings as in the proof of Lemma~\ref{step-p/thm} (cf.
Figure~\ref{onion-pf2/fig}), in order to push all the edges attached along the
$\alpha_i^s$ so that that they becomes attached to the last one exiting from $v_i$ (and
different from $e_i$). Then, the obvious $\phi$-invariant $F$-moves allow us to
change the resulting $T$ into the form depicted on the right side of
Figure~\ref{tripode3/fig}, with each $\alpha_i$ being now the shortest path between $v_i$
and $v_{i+1 \!\!\mod 3}$. Finally, we can get the configuration described in point~2 of
the statement and shown in the left side of Figure~\ref{tripode2/fig}, by further
$\phi$-invar\-iant slidings of the new edges attached along the actual $\alpha_i$ as
above.
Otherwise, if $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) > \mathop{\mathrm{len}}\nolimits(\alpha)/2$, then the same
argument as in the proof of Lemma~\ref{step-p/thm} still works here, to put
$\cup_i\,\phi^i(\alpha)$ in the form stated in point~1 of the statement.
\end{proof}
\begin{proposition}\label{onion-3/thm}
Let $\phi \in \cal A_{g,b}$ be a non-trivial automorphism, whose order $n = \mathop{\mathrm{ord}}\nolimits(\phi) > 1$ is
an odd multiple of $3$. Then, $\phi$ is $F$-equivalent to the composition $\tau \circ
\sigma$ of two automorphisms $\sigma, \tau \in \cal A_{g,b}$ such that $\mathop{\mathrm{ord}}\nolimits(\sigma) =
\mathop{\mathrm{ord}}\nolimits(\tau) = 2$. Moreover, both $\sigma$ and $\tau$ can be assumed to fix an edge and
reverse an invariant edge.
\end{proposition}
\begin{proof}
Given $\phi: \Gamma \to \Gamma$ as in the statement, the same recursive construction of
the proof of Proposition~\ref{onion-p/thm} will provide a sequence of automorphisms
$\phi_i: \Gamma_i \to \Gamma_i$ and a sequence of subgraphs $\Lambda_i \subset \Gamma_i$
with $i = 0, \dots, r$, satisfying the properties required in that proof, except for the
following facts:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[{\sl a\/})] $\Lambda_1$ is allowed to be a subgraph of the first
barycentric subdivision of $\Gamma_1$ instead that a subgraph of $\Gamma_1$ itself;
\@ifnextchar[{\item@}{\item@@}[{\sl b\/})] $U_i$ is required to have cardinality $n$ for $i = 1, \dots, r-1$
(Lemma~\ref{orders/thm} ensures that such cardinality is either $n$ or $n/3$).
\end{itemize}
In particular, all the $\phi_i$ are $F$-equivalent to $\phi$, and all the restrictions
$\phi_{i|\Lambda_i}$ admit a factorization $\tau_i \circ \sigma_i$ into two involutions of
$\Lambda_i$. Hence, $\phi$ turns out to be $F$-equivalent to $\phi_r = \tau_r \circ
\sigma_r$ (being $\Lambda_r = \Gamma_r$), which proves the proposition.
The starting step of the recursion is provided by Lemma~\ref{cycle-3/thm}. If the
situation described in point 1 of that lemma occurs, we define the automorphism $\phi_1:
\Gamma_1 \to \Gamma_1$, the subgraph $\Lambda_1 \subset \Gamma_1$, the two involutions
$\sigma_1,\tau_1: \Lambda_1 \to \Lambda_1$ and the vertex $u_1$, as in the proof of
Proposition~\ref{onion-p/thm}. Otherwise, the situation described in point 2 of
Lemma~\ref{cycle-3/thm} occurs. In this case, there exists a $\phi$-invariant disjoint
union of $n/3$ tripodes $T \sqcup \phi(T) \sqcup \dots \sqcup \phi^{n/3\,-\,1}(T) \subset
\Gamma$, with $\phi^{n/3}$ cyclically permuting the edges of the tripode $T$. Then, we put
$\Gamma_1 = \Gamma_0 = \Gamma$, $\phi_1 = \phi_0 = \phi$ and $\Lambda_1 = T' \sqcup
\phi(T') \sqcup \dots \sqcup \phi^{n/3\,-\,1}(T') \subset \Gamma_1$, where $T' \subset T$
is the star of the trivalent vertex of $T$ in the first barycentric subdivision. Moreover,
we denote by $v$ the trivalent vertex of $T'$ and by $u$ any free end of $T'$, and
we define $\sigma_1$ and $\tau_1$ by putting $\sigma_1(\phi^j(v)) = \phi^{-j}(v)$,
$\sigma_1(\phi^j(u)) = \phi^{-j}(u)$, $\tau_1(\phi^j(v)) = \phi^{1-j}(v)$ and
$\tau_1(\phi^j(u)) = \phi^{1-j}(u)$ for $j = 0, \dots, n-1$.
In order to conclude the starting step, it is enough to observe that: $\sigma_1$ and
$\tau_1$ are well-defined involutive automophisms of $\Lambda_1$, since they preserve the
adjacency of vertices, being $\phi^j(v)$ adjacent to $\phi^k(u)$ if and only if $j = k
\!\!\mod n/3$; $\tau_1 \circ \sigma_1$ coincides with the restriction
$\phi_{1|\Lambda_1}$; $\Delta_1 = \mathop{\mathrm{Cl}}\nolimits(\Gamma_1 - \Lambda_1)$ is a uni/trivalent graph;
$U_1 = \Delta_1 \cap \Lambda_1$ consists of the $\phi_1$-orbit of the univalent vertex $u$
of $\Delta_1$ and it has cardinality $n$; $u_1 = u$ is the only vertex of $U_1$ fixed by
$\sigma_1$.
Now, to realize the recursive step of the construction, assume we are given $\phi_{i-1}:
\Gamma_{i-1} \to \Gamma_{i-1}$, $u_{i-1} \in U_{i-1} \subset \Lambda_{i-1} \varsubsetneq
\Gamma_{i-1}$ and $\sigma_{i-1},\tau_{i-1}: \Lambda_{i-1} \to \Lambda_{i-1}$ with the
properties required in the proof of Proposition~\ref{onion-p/thm} for $i-1 < r$,
integrated by the points {\sl a\/} and {\sl b\/} said above. In particular, the
$\phi_{i-1}$-orbit $U_{i-1}$ has cardinality $n$. Therefore, since $\mathop{\mathrm{ord}}\nolimits(\phi_{i-1}) = n$
by Lemma~\ref{ordinv/thm}, also the restriction of $\phi_{i-1}$ to the uni/tri\-valent
graph $\Delta_{i-1} = \mathop{\mathrm{Cl}}\nolimits(\Gamma_{i-1} - \Lambda_{i-1})$ is an automorphism of the same
order $n$. Furthermore, the terminal edges of $\Delta_{i-1}$ ending at vertices in
$U_{i-1}$ are all distinct (see proof of Proposition~\ref{onion-p/thm}).
If $\Delta_{i-1}$ does not contain any path joining different terminal edges ending at
vertices in $U_{i-1}$, then we can terminate the recursion with the $i$-th step, in the
same way as in the proof of Proposition~\ref{onion-p/thm}.
Otherwise, if a path $\alpha \subset \Delta_{i-1}$ exists joining different terminal edges
of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$, then we can choose it to be minimal
(hence simple) and apply Lemma \ref{step-3/thm} to the restriction of $\phi_{i-1}$ to
$\Delta_{i-1}$ and to such a minimal $\alpha$, in order to get the structure described in
that statement for $\cup_j\, \phi_{i-1}^j(\alpha) \subset \Delta_{i-1}$ up to
$F$-equivalence. Such an $F$-equivalence only involves internal edges of $\Delta_{i-1}$,
hence it does not change the set of free ends $U_{i-1} \subset \Delta_{i-1}$ and the
restriction of $\phi_{i-1|\Delta_{i-1}}$. Therefore, it can be extended to an
$F$-equivalence of the whole $\phi_{i-1}$ on $\Gamma_{i-1}$, which leaves $\Lambda_{i-1}$
and the restriction $\phi_{i-1|\Lambda_{i-1}}$ unchanged.
As a result, we get a new uni/trivalent graph $\Gamma_i$, such that $\Lambda_{i-1} \subset
\Gamma_i$ and a new automorphism $\phi_i: \Gamma_i \to \Gamma_i$, which is $F$-equivalent
to $\phi_{i-1}$ and coincides with $\phi_{i-1}$ on $\Lambda_{i-1}$. We also get a new
minimal path $\alpha \subset \mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$, which joins different
$\phi_i$-equivalent terminal edges of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ ending at vertices
in $U_{i-1}$, such that $\cup_j\, \phi_i^j(\alpha)$ itself (no more up to
$F$-equivalence) has the structure stated in Lemma~\ref{step-3/thm}. Then, we proceed by
distinguishing the different situations described in points 1 to 3 of that lemma.
If the situation of point~1 occurs, then the $i$-th step is the same as in the proof of
Proposition~\ref{onion-p/thm}. In particular, if such step is not the concluding one, then
$U_i$ can be easily seen to have cardinality $n$.
In the case when the situation of point~2 occurs, we denote by $e \subset \Gamma_i$ the
terminal edge of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ whose free end is $u_{i-1}$. Then, we
choose $u_i$ to be the other end of $e$ and put $\Lambda_i = \Lambda_{i-1} \cup_j
\phi_i^j(e) \cup_j \phi_i^j(T)$, in such a way that $\Delta_i = \Gamma_i - \Lambda_i$ is a
uni/trivalent graph, which meets $\Lambda_i$ at the $\phi_i$-orbit $U_i$ of $u_i$.
Moreover, we define $\sigma_i$ and $\tau_i$ to be the unique automorphisms of $\Lambda_i$
extending $\sigma_{i-1}$ and $\tau_{i-1}$ respectively. The existence of such extensions
and all the required properties of them, in particular the equality $\tau_i \circ \sigma_i
= \phi_{i|\Lambda_i}$ and the fact that $u_i$ is the unique vertex of $U_i$ fixed by
$\sigma_i$, can be proved by the same argument as in the proof of Lemma~\ref{onion-p/thm}.
Finally, if the situation of point~3 occurs, we just put $\Lambda_i = \Lambda_{i-1}
\cup_j\, \phi^j_i(\hat T) = \Gamma_i$, where the last equality is due to the connectedness
of $\Gamma_i$, and define $\sigma_i$ and $\tau_i$ to be the unique extensions of
$\sigma_{i-1}$ and $\tau_{i_1}$ to $\Gamma_i$. These can be easily seen to verify all the
required properties for $i = r$, hence the recursion terminates with this step.
\end{proof}
\@startsection{section}{1}{\s@ctind{The case of order $2^m$%
\label{order-2m/sec}}
In order to deal with automorphisms $\phi: \Gamma \to \Gamma$ in $\cal A_{g,b}$ of order
$2^m$, we first prove that we can limit ourselves to consider the case when $\mathop{\mathrm{ord}}\nolimits_\phi(e)
= 2^m$ for every edge $e$ of $\Gamma$ (hence, according to Lemma \ref{orders/thm},
$\mathop{\mathrm{ord}}\nolimits_\phi(v) = 2^m$ for every vertex $v$ of $\Gamma$).
\begin{lemma}\label{reduction/thm}
Let $\phi:\Gamma \to \Gamma$ be an automorphism in $\cal A_{g,b}$ with $\mathop{\mathrm{ord}}\nolimits(\phi) = 2^m$, and
let $2^k = \min_{e \in \Gamma} \mathop{\mathrm{ord}}\nolimits_\phi(e)$, where $e$ varies among all the edges of
$\Gamma$. Then, up to $F$-equivalence and composition with elementary automorphisms,
$\phi$ can be reduced to an automorphism $\psi: \Gamma' \to \Gamma'$ such that $\mathop{\mathrm{ord}}\nolimits(\psi)
= 2^k$ and $\mathop{\mathrm{ord}}\nolimits_\psi(e) = 2^k$ for every edge $e$ of $\Gamma'$.
\end{lemma}
\begin{proof}
The proof is by induction on the pair $(m,n_m)$ with respect to the lexicographic order,
where $n_m$ is the number of the edges $e$ of $\Gamma$ such that $\mathop{\mathrm{ord}}\nolimits_\phi(e) = 2^m$.
The base of the induction is for $m = k$, and hence $\mathop{\mathrm{ord}}\nolimits_\phi(e) = 2^m$ for all the edges
$e$ of $\Gamma$, that is $n_m = 3g - 3 + 2b$.
If on the contrary $m > k$, then Lemma \ref{orders/thm} tells us that there
are three distinct edges $e_0,e_1$ and $e_2$ of $\Gamma$ sharing a trivalent vertex $v$ of
$\Gamma$, such that $\mathop{\mathrm{ord}}\nolimits_\phi(e_0) = \mathop{\mathrm{ord}}\nolimits_\phi(v) = 2^{m-1}$, $\mathop{\mathrm{ord}}\nolimits_\phi(e_1) =
\mathop{\mathrm{ord}}\nolimits_\phi(e_2) = 2^m$ and $e_2 = \phi^{2^{m - 1}}(e_1)$ (note that $e_1 \neq e_2$,
otherwise $e_1 = e_2$ would be a loop and its order should be $2^{m-1}$). Let $v_1$ and
$v_2$ the other ends of $e_1$ and $e_2$ respectively.
If $v_1$ and $v_2$ are (distinct) univalent vertices, then by composing with the terminal
switch $S_{e_1,e_2}$ we get an automorphism $\rho = S_{e_1,e_2} \circ \phi$ with either
$m(\rho) < m(\phi)$ or $m(\rho) = m(\phi)$ and $n_m(\rho) < n_m(\phi)$. In fact,
$\mathop{\mathrm{ord}}\nolimits_\rho(e) = \mathop{\mathrm{ord}}\nolimits_\phi(e)/2$ for all the edges $e$ in the $\phi$-orbit of $e_1$ (and
$e_2$), while $\mathop{\mathrm{ord}}\nolimits_\rho(e) = \mathop{\mathrm{ord}}\nolimits_\phi(e)$ for all the other edges of $\Gamma$.
If $v_1$ and $v_2$ are coinciding trivalent vertices, then the same argument as above
applies, but with $S_{e_1,e_2}$ an internal switch.
Finally, assume that $v_1$ and $v_2$ are distinct trivalent vertices. Then, we have the
situation depicted in the left side of Figure \ref{reduction/fig}, where
$\phi^{2^{m-1}}(a_i) = a_{3-i}$ and $\phi^{2^{m-1}}(b_i) = b_{3-i}$ for $i = 1,2$ (with
the four edges $a_1,a_2, b_1$ and $b_2$ not necessarily distinct).
\begin{Figure}[htb]{reduction/fig}
\fig{}{reduction.eps}
{}{Reducing $(m(\phi),n_m(\phi))$.}
\end{Figure}
In this case, we can perform the $\phi$-invariant move $F_{\cal T,\cal T'}$ acting on a
neighborhood of $e_1 \cup e_2$ as described in Figure \ref{reduction/fig}. This changes
$\phi$ into an automorphism $\rho$ with either $m(\rho) < m(\phi)$ or $m(\rho) = m(\phi)$
and $n_m(\rho) < n_m(\phi)$. In fact $F_{\cal T,\cal T'}$, replaces the $\phi$-orbit of
$e_1$ (and $e_2$) by two different $\rho$-orbits of $e'_1$ and $e'_2$ respectively, having
cardinality $2^{m-1}$. This completes proof of the inductive step.
\end{proof}
In the light of the previous lemma, the next two lemmas concern the special case of an
automorphism $\phi: \Gamma \to \Gamma$ with $\mathop{\mathrm{ord}}\nolimits(\phi) = n = 2^m$ and $\mathop{\mathrm{ord}}\nolimits_\phi(e) = n$
for every edge $e$ of $\Gamma$. We notice that, under the latter assumption, the situation
described in Lemma \ref{paths-doubled/thm} can occur only with $\phi^i(\alpha) =
\phi^j(\alpha)$, that is $i = j \!\!\mod n$.
\pagebreak
\begin{lemma}\label{cycle-2/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, with $\mathop{\mathrm{ord}}\nolimits(\phi) = n = 2^m$ and $\mathop{\mathrm{ord}}\nolimits_\phi(e) = n$ for every
edge $e$ of $\,\Gamma$.\break If $\alpha \subset \Gamma$ is a (simple) path of minimal
length among all paths joining any two distinct $\phi$-equivalent vertices, then $\cup_i\,
\phi^i(\alpha)$ is one of the following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] a disjoint union $C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$
of $n/\ell$ simple cycles (possibly a single one, for $\ell = n$), each given by a
concatenation of images of $\alpha$, having the same structure as described in Lemma
\ref{cycle-p/thm}; in this case, up to $F$-equivalence we can assume $\alpha$ to consist
of a single edge $a$ (see Figure \ref{cycle-2/fig}, where $\ell$ is a power of $2$,
$a_i$ stands for $\phi^{is}(a)$ and a similar notation is adopted for the vertices $v_i$
as well);
\begin{Figure}[htb]{cycle-2/fig}
\fig{}{cycle-2.eps}
{}{The form of the cycle $C$ in Lemma~\ref{cycle-2/thm}.}
\end{Figure}
\@ifnextchar[{\item@}{\item@@}[2)] a disjoint union $D \sqcup \phi(D) \sqcup \dots \sqcup \phi^{n/2\,-\,1}(D)$ of
$n/2$ ``diagonal'' edges (possibly a single one, for $n = 2$), where $D = \alpha$ and
$\phi^{n/2}(D) = \bar D$ (see Figure~\ref{diagonal1/fig}, where the vertices $v_0$ and
$v_1$ are switched by $\phi^{n/2}$, and they can be either both univalent as on the left
side or both trivalent as on the right side).
\end{itemize}
\end{lemma}
\begin{Figure}[htb]{diagonal1/fig}
\vskip-6pt
\fig{}{diagonal1.eps}
{}{The ``diagonal'' edge $D$ in Lemma~\ref{cycle-2/thm}.}
\end{Figure}
\begin{proof}
Let $\alpha \subset \Gamma$ be an arbitrary path of minimal length among all paths joining
any two distinct $\phi$-equivalent vertices and let $v \neq v'$ be its ends. For any $i
\neq j \!\!\mod n$, let us consider the two paths $\phi^i(\alpha)$ and $\phi^j(\alpha)$,
and the four different situations described in Lemmas~\ref{paths-disjoint/thm} to
\ref{paths-doubled/thm}, which cover all the possibilities, thanks to the
$\phi$-equivalence of $v$ and $v'$.
The situation of Lemma~\ref{paths-doubled/thm} cannot occur, due to the choice of $i$ and
$j$, while any of the three remaining situations may occur. In particular,
$\phi^i(\alpha)$ and $\phi^j(\alpha)$ can have non-empty intersection both as in
Lemma~\ref{paths-diagonal/thm} and as in Lemma~\ref{paths-adjacent/thm}.
Let us assume first that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ are as in
Lemma~\ref{paths-diagonal/thm}, i.e. they share both ends, with $\phi^i(v) = \phi^j(v')$
and $\phi^j(v) = \phi^i(v')$. In particular, $\phi^{j-i}$ switches $v$ and $v'$, and $j -
i = n/2 \!\!\mod n$, since $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n$.
The global minimality of $\alpha$ ensures that either $\alpha$ is a single ``diagonal''
edge $D$ with $\phi^{n/2}(D) = \bar D$, or $\alpha$ and $\phi^{n/2}(\alpha)$ only meet at
\pagebreak
their common ends $v$ and $v'$. Moreover, $\alpha$ and $\phi^{n/2}(\alpha)$ are disjoint
from any other $\phi^k(\alpha)$ with $k \neq 0,n/2 \!\!\mod n$, thanks to
Lemma~\ref{paths-disjoint/thm}. If $\alpha$ is a single ``diagonal'' edge $D$, we have
that $\cup_i\, \phi^i(\alpha)$ is as in point 2 of the statement. Otherwise,
$\cup_i\,\phi^i(\alpha)$ consists in the disjoint union $C \sqcup \phi(C) \sqcup \dots
\sqcup \phi^{n/2 - 1}(C)$ of $n/2$ simple cycles (possibly a single one, for $n = 2$),
with $C = \alpha \cup \phi^{n/2}(\alpha)$. Then, arguing as in the proof of
Lemma~\ref{cycle-p/thm}, the length of $\alpha$ (and that of its images as well) may be
reduced to 1, which gives a special case of the situation described in point 1 of the
statement.
At this point, we are left with the case when $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can
only meet as in Lemma~\ref{paths-adjacent/thm}. In this case, denoting by $\phi^i(u)$ with
$u \in \alpha$ the common end of $\delta_{i,1}$ and $\delta_{j,k}$, we have that
$\phi^j(u) \in \delta_{j,k}$ since $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) \geq
\mathop{\mathrm{len}}\nolimits(\alpha)/2$. Actually, $\phi^j(u)$ has to coincide with either $\phi^j(v')$ or
$\phi^i(u)$, otherwise the global minimality of $\alpha$ would be contradicted. However,
$\phi^j(u) = \phi^i(u)$ is not possible because this would imply that $\mathop{\mathrm{ord}}\nolimits_\phi(u) < n$.
Thus, $\phi^j(u) = \phi^j(v')$ and we can conclude that the set $\cup_i\, \phi^i(\alpha)$
has the structure described in point 1 of the statement, by arguing as in the proof of
Lemma~\ref{cycle-p/thm}.
\end{proof}
\begin{lemma}\label{step-2/thm}
Let $\phi: \Gamma \to \Gamma$ be an automorphism of a (possibly disconnected)
uni/trivalent graph $\Gamma$, with $\mathop{\mathrm{ord}}\nolimits(\phi) = n = 2^m$ and $\mathop{\mathrm{ord}}\nolimits_\phi(e) = n$ for every
edge $e$ of $\,\Gamma$.\break If $\alpha \subset \Gamma$ is a (simple) path of minimal
length among all paths joining any two distinct terminal edges in a given $\phi$-orbit,
then up to $F$-equivalence we can assume $\cup_i\,\phi^i(\alpha)$ to be one of the
following:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[1)] of a disjoint union
$C \sqcup \phi(C) \sqcup \dots \sqcup \phi^{n/\ell - 1}(C)$ of $n/\ell$ simple cycles
(possibly a single one for $\ell = n$), each given by a concatenation of images of
$\alpha$, having the same structure as described in Lemma~\ref{step-p/thm}, with $\alpha$
consisting of either one edge $a$ or two edges $a$ and $b$ (see Figure~\ref{onion-2/fig},
where $\ell$ is a power of $2$, $a_i$ and $b_i$ stand for $\phi^{is}(a)$ and
$\phi^{is}(b)$ respectively, and a similar notation is adopted for the terminal edges
$e_i$'s and the vertices $v_i$'s as well);
\begin{Figure}[htb]{onion-2/fig}
\vskip-6pt
\fig{}{onion-2.eps}
{}{The form of the cycle $C$ in Lemma~\ref{step-2/thm}.}
\end{Figure}
\@ifnextchar[{\item@}{\item@@}[2)] a disjoint union $D \sqcup \phi(D) \sqcup \dots \sqcup \phi^{n/2\,-\,1}(D)$ of
$n/2$ ``diagonal'' edges (possibly a single one, for $n = 2$), where $D = \alpha$ and
$\phi^{n/2}(D) = \bar D$ (see Figure~\ref{diagonal2/fig}, where the edges $e_0$ and
$e_1$ are switched by $\phi^{n/2}$).
\end{itemize}
\end{lemma}
\begin{Figure}[htb]{diagonal2/fig}
\vskip6pt
\fig{}{diagonal2.eps}
{}{The ``diagonal'' edge $D$ in Lemma~\ref{step-2/thm}.}
\end{Figure}
\begin{proof}
Let $\alpha$ be a path as in the statement, and let $e \neq e'$ be the terminal edges it
joins. Then the ends of $\alpha$ coincide with the unique trivalent ends $v$ and $v'$ of
$e$ and $e'$ respectively. Notice that $v \neq v'$ since otherwise $\mathop{\mathrm{ord}}\nolimits_\phi(v)$ would be
less than $n$.
For $i \neq j \!\!\mod n$, the two paths $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can meet
both as in Lemma~\ref{paths-adjacent/thm} and as in Lemma~\ref{paths-diagonal/thm}, but
not as in Lemma~\ref{paths-doubled/thm}.
Let us assume first that $\phi^i(\alpha)$ and $\phi^j(\alpha)$ are as in
Lemma~\ref{paths-diagonal/thm}, i.e. they share both ends, with $\phi^i(v) = \phi^j(v')$
and $\phi^j(v) = \phi^i(v')$, and their union $\phi^i(\alpha) \cup \phi^j(\alpha)$ looks
like in Figure~\ref{paths-diagonal/fig}. In particular, $\phi^{j-i}$ switches $v$ and
$v'$, and then $j - i = n/2 \mod n$, since $\mathop{\mathrm{ord}}\nolimits_\phi(v) = n$. Moreover,
Lemma~\ref{paths-disjoint/thm} ensures that $\phi^i(\alpha) \cup \phi^j(\alpha)$ is
disjoint from any other path $\phi^k(\alpha)$ with $k \neq i,j \!\!\mod n$.
Hence, $\cup_i\,\phi^i(\alpha)$ is a disjoint union of $n/2$ pairs of ``diagonal'' paths,
each pair being given by $\phi^i(\alpha)$ and $\phi^{i+n/2}(\alpha)$ for some $i = 0,
\dots, n/2-1$. Figure~\ref{onion-2-pf1/fig} shows the two possible forms of
$\phi^i(\alpha) \cup \phi^{i+n/2}(\alpha)$, depending on the fact that $\alpha$ and
$\phi^{n/2}(\alpha)$ start and end with non-trivial common subpaths (left side) or not
(right side). Here, the subscripts denote the images under the corresponding power of
$\phi$, and apart from the edges $e_i$ and $e_{i+n/2}$ each arc represents a path of
edges. We notice that in both cases, if the length of the path $\alpha$ is even, then the
central vertices of the paths $\alpha_i$ and $\alpha_{i+n/2}$ have to be distinct (that
is, they have to belong to non-common subarcs), otherwise the order of such vertices would
be $n/2$.
\begin{Figure}[htb]{onion-2-pf1/fig}
\fig{}{onion-2-pf1.eps}
{}{The starting form of $\phi^i(\alpha) \cup \phi^{i+n/2}(\alpha)$ in the proof of
Lemma~\ref{step-2/thm}.}
\end{Figure}
\begin{Figure}[b]{onion-2-pf2/fig}
\fig{}{onion-2-pf2.eps}
{}{Simplifying the configuration on the left side of Figure~\ref{onion-2-pf1/fig}.}
\end{Figure}
The configuration on the left side of Figure~\ref{onion-2-pf1/fig} can be simplified by a
sequence of symmetric pairs of $\phi$-invariant slidings as indicated in
Figure~\ref{onion-2-pf2/fig} (cf. Figure~\ref{onion-pf2/fig}), in order to reduce all the
arcs to single edges. After that, all the resulting bigons, except the
central one if they are odd in number, can be modified in pairs as in
Figure~\ref{tripode3/fig} and then slided out of $\phi^i(\alpha) \cup
\phi^{i+n/2}(\alpha)$ as above. Then, in the case of an even number of bigons we end up
with the situation described in point 1 of the statement. Otherwise, we are left with one
bigon as shown on the left side of Figure~\ref{onion-2-pf3/fig}, and we can perform the
\pagebreak
$F$-move described in that figure, to get the configuration on the right side of the same
Figure~\ref{onion-2-pf3/fig}. This is the special case for $\ell = 2$ of the configuration
on the right side of Figure~\ref{onion-2/fig} considered in point 1 of the statement.
\begin{Figure}[htb]{onion-2-pf3/fig}
\vskip-6pt
\fig{}{onion-2-pf3.eps}
{}{Simplified version of the configuration on the left side of
Figure~\ref{onion-2-pf1/fig}.}
\end{Figure}
Now, look at the configuration on the right side of Figure~\ref{onion-2-pf1/fig}. If
$\mathop{\mathrm{len}}\nolimits(\alpha) = 1$, that is $\phi^i(\alpha) \cup \phi^{i+n/2}(\alpha)$ consists of a
single bigon, then we just have the special case for $\ell = 2$ of the configuration on
the left side of Figure~\ref{onion-2/fig} considered in point 1 of the statement. If
$\mathop{\mathrm{len}}\nolimits(\alpha) > 1$, then we can simplify the configuration by the same argument as above,
but sliding everything to the most external loops adjacent to $e_i$ and $e_{i + n/2}$.
The result is shown on the left side of Figure~\ref{onion-2-pf4/fig}. We can change it
into the configuration on the right side of the figure, by a $\phi$-invariant $F$-move on
the central edge, and then to the configuration on the right side of
Figure~\ref{onion-2-pf3/fig}, by a further $\phi$-invariant sliding (reducing the length
of $\alpha$ to 2). So, we get once again a special case of the situation described in
point 1 of the statement.
\begin{Figure}[htb]{onion-2-pf4/fig}
\fig{}{onion-2-pf4.eps}
{}{Simplified version of the configuration on the right side of
Figure~\ref{onion-2-pf1/fig}.}
\end{Figure}
At this point, we are left with the case when $\phi^i(\alpha)$ and $\phi^j(\alpha)$ can
only meet as in Lemma~\ref{paths-adjacent/thm}. In this case, then the fact that $n$ is
even ensures that their common end $\phi^j(v) = \phi^i(v')$ cannot be shared by any other
$\phi^h(\alpha)$ and that $\mathop{\mathrm{len}}\nolimits(\delta_{i,1}) = \mathop{\mathrm{len}}\nolimits(\delta_{j,k}) > \mathop{\mathrm{len}}\nolimits(\alpha)/2$ (cf.
Figure \ref{paths-adjacent/fig}). Therefore, the same arguments as in
Lemma~\ref{cycle-p/thm} ensure that the situation can be reduced to the one described in
point 1 of the statement.
\end{proof}
\begin{proposition}\label{onion-2/thm}
Let $\phi: \Gamma \to \Gamma$ be a non-trivial automorphism in $\cal A_{g,b}$, with
$\mathop{\mathrm{ord}}\nolimits(\phi) = n = 2^m$. Then, up to $F$-equivalence and composition with elementary
automorphisms, we can reduce $\phi$ to the composition $\tau \circ \sigma$ of two
automorphisms $\sigma, \tau \in \cal A_{g,b}$ such that $\mathop{\mathrm{ord}}\nolimits(\sigma) = \mathop{\mathrm{ord}}\nolimits(\tau) = 2$.
Moreover, $\sigma$ can be assumed to fix an edge and $\tau$ can be assumed to reverse an
invariant edge.
\end{proposition}
\begin{proof}
Given $\phi: \Gamma \to \Gamma$ as in the statement, thanks to Lemma~\ref{reduction/thm}
we can as\-sume that $\mathop{\mathrm{ord}}\nolimits_\phi(e) = n$ for every edge $e$ of $\,\Gamma$. Then, the same
recursive construction of the proof of Proposition~\ref{onion-p/thm} will provide a
sequence of automorphisms $\phi_i: \Gamma_i \to \Gamma_i$ and a sequence of subgraphs
$\Lambda_i \subset \Gamma_i$ with $i = 0, \dots, r$, satisfying the properties required in
that proof, except for the following facts:
\begin{itemize}
\@ifnextchar[{\item@}{\item@@}[{\sl a\/})] $\Lambda_1$ is allowed to be a subgraph of the second barycentric
subdivision of $\Gamma_1$ instead that a subgraph of $\Gamma_1$ itself;
\pagebreak
\@ifnextchar[{\item@}{\item@@}[{\sl b\/})] $\mathop{\mathrm{ord}}\nolimits_{\phi_i}(e) = n$ for every edge $e$ of $\Gamma_i$ and every $i =
1, \dots, r$;
\@ifnextchar[{\item@}{\item@@}[{\sl c})] $\sigma_i$ fixes exactly two vertices $u_i$ and $\phi_i^{n/2}(u_i)$ in
$U_i$, for every $i = 0, \dots, r-1$;
\@ifnextchar[{\item@}{\item@@}[{\sl d})] $\sigma_i$ fixes two edges of $\Gamma_i$, but it is not required to
reverse any invariant edge, for every $2 \leq i \leq r$;
\@ifnextchar[{\item@}{\item@@}[{\sl e})] the $\tau_i$'s are not required to fix or reverse any edge, except for
$\tau_r$ (the last one) that has to reverse an invariant edge.
\end{itemize}
In particular, all the $\phi_i$'s are $F$-equivalent to $\phi$, and all the restrictions
$\phi_{i|\Lambda_i}$ admit a factorization $\tau_i \circ \sigma_i$ into two involutions of
$\Lambda_i$. Hence, $\phi$ turns out to be $F$-equivalent to $\phi_r = \tau_r \circ
\sigma_r$ (being $\Lambda_r = \Gamma_r$), which proves the proposition.
\vskip\smallskipamount
The starting step of the recursion is provided by Lemma~\ref{cycle-2/thm}. If the
situation described in point 1 of that lemma occurs, we define the automorphism $\phi_1:
\Gamma_1 \to \Gamma_1$, the subgraph $\Lambda_1 \subset \Gamma_1$, the two involutions
$\sigma_1,\tau_1: \Lambda_1 \to \Lambda_1$ and the vertex $u_1$, as in the proof of
Proposition~\ref{onion-p/thm}. Otherwise, the situation described in point 2 of
Lemma~\ref{cycle-2/thm} occurs. In this case, there exists a $\phi$-invariant disjoint
union of $n/2$ ``diagonal'' edges $D \sqcup \phi(D) \sqcup \dots \sqcup
\phi^{n/2\,-\,1}(D) \subset \Gamma$, with $\phi^{n/2}(D) = \bar D$. Then, we put $\Gamma_1
= \Gamma_0 = \Gamma$, $\phi_1 = \phi_0 = \phi$ and $\Lambda_1 = D' \sqcup \phi(D') \sqcup
\dots \sqcup \phi^{n/2\,-\,1}(D') \subset \Gamma_1$, where $D' \subset D$ is the start of
the barycenter of $D$ in the second barycentric subdivision of $\Gamma$. Moreover, we
denote by $u$ any one of the two ends of $D'$, and we define $\sigma_1$ and $\tau_1$ by
putting $\sigma_1(\phi^j(u)) = \phi^{-j}(u)$ and $\tau_1(\phi^j(u)) = \phi^{1-j}(u)$ for
$j = 0, \dots, n-1$.
In order to conclude the starting step, it is enough to observe that: $\mathop{\mathrm{ord}}\nolimits_{\phi_1}(e) =
n$ for every edge $e$ of $\Gamma_1$; $\sigma_1$ and $\tau_1$ are well-defined involutive
automophisms of $\Lambda_1$, since they preserve the adjacency of vertices, being
$\phi^j(u)$ adjacent to $\phi^k(u)$ if and only if $j = k \!\!\mod n/2$; $\tau_1 \circ
\sigma_1$ coincides with the restriction $\phi_{1|\Lambda_1}$; $\Delta_1 = \mathop{\mathrm{Cl}}\nolimits(\Gamma_1 -
\Lambda_1)$ is a uni/trivalent graph; $U_1 = \Delta_1 \cap \Lambda_1$ consists of the
$\phi_1$-orbit of the univalent vertex $u$ of $\Delta_1$ and it has cardinality $n$; $u_1
= u$ and $\phi^{n/2} (u_1)$ are the only vertices of $U_1$ fixed by $\sigma_1$.
Now, to realize the recursive step of the construction, assume we are given $\phi_{i-1}:
\Gamma_{i-1} \to \Gamma_{i-1}$, $u_{i-1} \in U_{i-1} \subset \Lambda_{i-1} \varsubsetneq
\Gamma_{i-1}$ and $\sigma_{i-1},\tau_{i-1}: \Lambda_{i-1} \to \Lambda_{i-1}$ with the
properties required in the proof of Proposition~\ref{onion-p/thm} for $i-1 < r$,
integrated by the points {\sl a\/} to {\sl e\/} said above. In particular, the restriction
of $\phi_{i-1}$ to the uni/tri\-valent graph $\Delta_{i-1} = \mathop{\mathrm{Cl}}\nolimits(\Gamma_{i-1} -
\Lambda_{i-1})$ is an automorphism of order $n$, and the order of each edge of
$\Delta_{i-1}$ with respect to it is $n$ as well.
If two of the terminal edges of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$ coincide,
then there are $n/2$ such terminal edges and each of them is a ``diagonal'' edge reversed
by $\phi_{i-1}^{n/2}$. In this case, denoting by $D$ the ``diagonal'' edge between
$u_{i-1}$ and $\phi_{i-1}^{n/2}(u_{i-1})$, we put $\Gamma_i = \Gamma_{i-1}$, $\phi_i =
\phi_{i-1}$ and $\Lambda_i = \Lambda_{i-1} \cup_j\,\phi^j_{i-1}(D)$, where the last
equality is due to the connectedness of $\Gamma_i$. Moreover, we define $\sigma_i$ and
$\tau_i$ to be the unique extensions of $\sigma_{i-1}$ and $\tau_{i-1}$ to $\Gamma_i$.
These can be easily seen to exist and to verify all the required properties for $i = r$,
except for the fact that $\tau_r$ reverses some invariant edge, which\break will be proved
at the end of the proof. Hence, this step terminates the recursion.
Then, let the terminal edges of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$ be all
distinct.
If $\Delta_{i-1}$ does not contain any path joining two different such terminal edges,
then we can terminate the recursion with the $i$-th step, in the same way as in the proof
of Proposition~\ref{onion-p/thm}. Once again, we postpone at the end of the proof the
verification that $\tau_r$ reverses some invariant edge.
Otherwise, if a path $\alpha \subset \Delta_{i-1}$ exists joining different terminal edges
of $\Delta_{i-1}$ ending at vertices in $U_{i-1}$, then we can choose it to be minimal
(hence simple) and apply Lemma \ref{step-2/thm} to the restriction of $\phi_{i-1}$ to
$\Delta_{i-1}$ and to such a minimal $\alpha$, in order to get the structure described in
that statement for $\cup_j\, \phi_{i-1}^j(\alpha) \subset \Delta_{i-1}$ up to
$F$-equivalence. Such an $F$-equivalence only involves internal edges of $\Delta_{i-1}$,
hence it does not change the set of free ends $U_{i-1} \subset \Delta_{i-1}$ and the
restriction of $\phi_{i-1|\Delta_{i-1}}$. Therefore, it can be extended to an
$F$-equivalence of the whole $\phi_{i-1}$ on $\Gamma_{i-1}$, which leaves $\Lambda_{i-1}$
and the restriction $\phi_{i-1|\Lambda_{i-1}}$ unchanged.
As a result, we get a new uni/trivalent graph $\Gamma_i$, such that $\Lambda_{i-1} \subset
\Gamma_i$ and a new automorphism $\phi_i: \Gamma_i \to \Gamma_i$, which is $F$-equivalent
to $\phi_{i-1}$ and coincides with $\phi_{i-1}$ on $\Lambda_{i-1}$. We also get a new
minimal path $\alpha \subset \mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$, which joins different
$\phi_i$-equivalent terminal edges of $\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ ending at vertices
in $U_{i-1}$, such that $\cup_j\, \phi_i^j(\alpha)$ itself (no more up to
$F$-equivalence) has the structure stated in Lemma~\ref{step-2/thm}. Then, we proceed by
distinguishing the different situations described in points 1 and 2 of that lemma.
If the situation of point~1 occurs, then the $i$-th step is the same as in the proof of
Proposition~\ref{onion-p/thm}, but now $\sigma_i$ fixes two edges and $\tau_i$ reverses
two invariant edges.
In the case when the situation of point~2 occurs, we denote by $e$ the terminal edge of
$\mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_{i-1})$ whose free end is $u_{i-1}$. Then, we choose $u_i$ to be
the other end of $e$ and put $\Lambda_i = \Lambda_{i-1} \cup_j\, \phi_i^j(e) \cup_j\,
\phi_i^j(D)$, in such a way that $\Delta_i = \mathop{\mathrm{Cl}}\nolimits(\Gamma_i - \Lambda_i)$ is a uni/trivalent
graph, which meets $\Lambda_i$ at the $\phi_i$-orbit $U_i$ of $u_i$. Moreover, we define
$\sigma_i$ and $\tau_i$ to be the unique automorphisms of $\Lambda_i$ extending
$\sigma_{i-1}$ and $\tau_{i-1}$ respectively. To see that such extensions exist, we first
define $\sigma_i(\phi_i^j)(u_i) = \phi_i^{-j}(u_i)$ and $\tau_i(\phi_i^j)(u_i) =
\phi_i^{1-j}(u_i)$, for every $j = 0, \dots, n-1$, and then we verify that these
definitions, together with the previous definitions of $\sigma_{i-1}$ and $\tau_{i-1}$,
preserve adjacency of vertices in $\Lambda_i$. Notice that $\tau_i \circ \sigma_i$
trivially coincides with $\phi_{i|\Lambda_i}$. Furthermore, $\sigma_i$ fixes only the
vertices $u_i$ and $\phi_i^{n/2}(u_i)$ in the $\phi_i$-orbit $U_i$, and it fixes the edges
$e$ and $\phi^{n/2}(e)$. On the contrary, $\tau_i$ neither fixes nor reverses any edge in
the $\phi_i$-orbits of $e$ and $D$. This conclude the recursive step.
Finally, we observe that the situation of point 1 of Lemma~\ref{step-2/thm} occurs in at
least one of the recursive steps due to the connectedness of $\Gamma$, and from that
step on all the $\tau_i$ reverse some invariant edge. In particular, this is true for
$\tau_r$.
\end{proof}
At this point, taking into account Propositions~\ref{onion-p/thm}, \ref{onion-3/thm} and
\ref{onion-2/thm}, we are left to show that any automorphism in $\cal A_{g,b}$ belongs to
$\cal E_{g,b}$, in the special case when it has order 2 and fixes or reverses some edge. This
is the content of our last proposition.
\begin{proposition}\label{automorphism-2/thm}
Let $\phi:\Gamma \to \Gamma$ be an automorphism in $\cal A_{g,b}$ of order $2$, which either
fixes an edge or reverses an invariant edge. Then, up to $F$-equivalence and composition
with elementary switches, we can reduce $\phi$ to the identity.
\end{proposition}
\begin{proof}
If $\phi$ fixes an edge $e$, then Lemma~\ref{reduction/thm} ensures that up to
$F$-equivalence and composition with elementary automorphisms, $\phi$ can be reduced to an
automorphism $\psi: \Gamma' \to \Gamma'$ such that $\mathop{\mathrm{ord}}\nolimits(\psi) = 1$, that is $\psi$
reduces to the identity. On the other hand, if $\phi$ reverses an invariant edge $e$, we
can perform a $\phi$-invariant $F$-move $F_{e,e'}$, in order to get a new automorphism
$\phi': \Gamma' \to \Gamma'$, which fixes the edge $e'$ of $\Gamma'$.
\end{proof}
\newpage
\bibliographystyle{amsplain}
|
\section{Introduction}\label{intro}
Let $S=S_{g,n}$ be an oriented surface of genus $g\geq 0$ with $n\geq 0$ punctures. We assume that $3g-3+n > 0$.
The Teichm\"uller space
$\mathcal{T}(S)$ of $S$ is the space of complex structures (or, equivalently, of complete and finite-area hyperbolic structures) on $S$
up to equivalence, where two such structures $X$ and $Y$ are considered as equivalent if there is a conformal map (respectively, an isometry) $h:(S,X) \to (S,Y)$ which is
homotopic to the identity map of $S$.
There are several natural metrics on $\mathcal{T}(S)$; some of them are Riemannian and others are only Finsler. Some metrics on Teichm\"uller space (e.g. the Weil-Petersson metric, and the Teichm\"uller metric) are strongly related to the complex structure of that space. A Finsler metric is a length metrics where the distance between two points is defined by minimizing lengths of peicewise $C^1$ paths joining them and where the length of a path is computed by integrating norms of tangent vectors. The fact that a metric is Finsler but not Riemannian means that the norm function on each tangent space is not associated to a scalar product. In this paper, we consider the Teichm\"uller metric and Thurston's asymmetric metric. Both metrics are Finsler, and in the case of Thurston's asymmetric metric, the Finsler norm not only is non-Riemannian but it is even not symmetric. These metrics are defined in terms of distances between geometric structures on the surface (conformal structures, in the case of the Techm\"uller metric, and hyperbolic structures, in the case of Thurston's asymmetric metric). The question of comparing the various metrics on $\mathcal{T}(S)$ is a natural one, and it was also suggested by Thurston in \cite{Thurston}; see also \cite{PT}.
We shall recall below the definition of Thurston's asymmetric metric.
Thurston formulated several ideas concerning this asymmetric metric, and we shall review some of them below. He proved some of the major results and he outlined some other results in his paper \cite{Thurston}. After Thurston's paper was circulated, questions on the limiting behavior of stretch lines and of anti-stretch lines were considered, see \cite{P1}, \cite{PT}, \cite{T3} and \cite{T2}. Stretch lines are some special geodesics of that metric, and anti-stretch lines are stretch lines traversed in the opposite directions; we note that since the metric is not symmetric, a geodesic traversed in the opposite direction is not necessarily a geodesic. There were also generalizations of Thurston's asymmetric metric to the case of surfaces with boundary, see \cite{LSPT1} and \cite{LSPT2}. A renewal of interest in this metric has emerged recently, see e.g. the papers \cite{Walsh}, \cite{LRT} and \cite{LPST}. It became also clear that among the known metrics on Teichm\"uller space, the Thurston metric is the one that has an interesting analogue on outer space, see \cite{CV}. Thus, there are very good reasons to study Thurston's asymmetric metric.
Some analogies and some differences between Thurston's asymmetric metric and Teichm\"uller's metric have described in the paper \cite{PT}. In the present paper, we point out new analogies that concern the Finsler character of these metrics.
Given a set $X$, a function $d:X\to X$ is said to be a \emph{weak metric} on $X$ if it satisfies all the axioms of a distance function except the symmetry axiom. The weak metric $d$ is said to be \emph{asymmetric} if it is strictly weak, that is if there exist two points $x$ and $y$ in $X$ such that $d(x,y)\not=d(y,x)$.
A related notion is that of a \emph{weak norm} on a vector space $E$. This is a function $p:E\to \mathbb{R}$ that is nonnegative, convex and positively homegeneous. In other words, $p$ satisfies the following:
\begin{enumerate}\label{weak-norm}
\item $p(x)\geq 0$ for all $x\in E$;
\item \label{wn1} $p(\lambda x)=\lambda p(x)$ for all $x\in E$ and for all $\lambda\geq 0$;
\item \label{wn2} $p(x_1+x_2) \leq p(x_1)+ p(x_2)$ for all $x_1,x_2 \in E$.
\end{enumerate}
In this paper, to simplify terminology, we shall use in general the term \emph{norm} to denote a weak norm a,d the terms \emph{metric} to denote a weak metric.
Let $X$ and $Y$ be now two complete hyperbolic metrics on $S$.
In the paper \cite{Thurston}, Thurston defined an asymmetric metric $d_L$ on $\mathcal{T}(S)$ by setting
\begin{equation} \label{eq:L}
d_L(X,Y)=\inf_{f} \log L_f(X,Y),
\end{equation}
where the infimum is taken is over all homeomorphsims $f:X \to Y$ homotopic to the indentity map of $S$, and where $L_f(X,Y)$ is the Lipschitz constant of $f$, that is,
\begin{displaymath}\label{Lip}
\hbox{Lip}(f)=\sup_{x\neq y\in S}\frac{d_{Y}\big{(}f(x),f(y)\big{)}}{d_{X}\big{(}x,y\big{)}}.
\end{displaymath}
We shall call this weak metric Thurston's asymmetric metric or, for short, Thurston's metric.
In the same paper, Thurston proved that there is a (non-necessarily unique) extremal Lipschitz homeomorphsim that realizes the infimum in (\ref{eq:L}). He also proved that we have the following formula for the asymmetric metric:
\begin{equation} \label{eq:LL}d_L(X,Y)=\log \sup_{\gamma} \frac{\ell_Y(\gamma)}{\ell_X(\gamma)},
\end{equation}
where $\ell_X(\gamma)$ denotes the hyperbolic length of $\gamma$ with respect to the metric $X$ and $\gamma$ ranges over all essential simple closed curves on $S$.
Furthermore, Thurston proved that the asymmetric metric defined in (\ref{eq:L}) is Finsler, that is, it is a length metric which is defined by integrating a weak norm on the tangent bundle of $\mathcal{T}(S)$ along paths in Teichm\"uller space, and taking the minimum lengths over all peicewise $C^1$-paths. Thurston gave an explicit formula for the weak norm of a tangent vector $V$ at a point $X$ in $\mathcal{T}(S)$, namely,
\begin{equation}\label{eq:norm-L}
\Vert V \Vert_L= \sup_{\lambda \in \mathcal{ML}} \frac{d \ell_{\lambda}(V)}{\ell_{\lambda}(X)}.
\end{equation}
Here, $\mathcal{ML}=\mathcal{ML}(S)$ is the space of measured laminations on $S$, $\ell_{\lambda}:\mathcal{T}(S)\to \mathbb{R}$ is the length function on Teichm\"uller space associated to the measured lamination $\lambda$ and $d \ell_{\lambda}$ is the differential of the function $\ell_{\lambda}(X)$ at the point $X\in \mathcal{T}(S)$.
For $X$ in $\mathcal{T}(S)$ and $\lambda$ in $\mathcal{ML}$, we shall use the notation $\ell_\lambda(X)$ or $\ell_X(\lambda)$ to denote the $X$-length of $\lambda$, depending on whether we consider the length function as a function on Teichm\"uller space or on measured lamination space.
We shall present an analogue of Formula (\ref{eq:norm-L})
for the Teichm\"uller metric that is expressed in terms of extremal length, namely, we show that the Finsler weak norm associated to the Teichm\"uller distance is given by the following formula:
\begin{equation}\label{eq:norm-T}
\Vert V\Vert_E= \sup_{\lambda \in \mathcal{MF}} \frac{d \mathrm{Ext}^{1/2}_{\lambda}(V)}{ \mathrm{Ext}^{1/2}_{\lambda}(X)} .
\end{equation}
Here, $X$ is a conformal structure on $S$, considered as a point in Teichm\"uller space, $V$ is a tangent vector at the point $X$, $\mathcal{MF}=\mathcal{MF}(S)$ is the space of measured laminations on $S$, $\mathrm{Ext} _{\lambda}:\mathcal{T} \to \mathbb{R}$ is the extremal length function associated to the measured foliation $\lambda$ and $d\mathrm{Ext} _{\lambda}$ is the differential of that function at $X$.
There is a more geometric version of (\ref{eq:norm-T}) in which the tangent vector $V$ is interpreted as a Beltrami differential (see Corollary \ref{coro:teich} below).
The following table summarizes some analogies between notions and results associated to Thurston's asymmetric metric and those associated to the
Teichm\"uller metric.
\bigskip
\begin{center}
\begin{tabular}{|c | l | l |}
\hline
{} & Thurston's metric & Teichm\"uller metric \\
\hline
(i) & Stretch maps & Teichm\"uller extremal maps \\
\hline
(ii) & Stretch lines & Teichm\"uller lines \\
\hline
(iii) & $d_L(X,Y)=\displaystyle \log \inf_f L(f)$ & $ d_T(X,Y)= \displaystyle\log \inf_f K(f) $\\
\hline
(iv) & $d_L(X,Y)=\log \displaystyle \sup_{\gamma} \displaystyle \frac{\ell_\gamma(Y)}{\ell_\gamma(X)}$ & $d_T(X,Y)= \log \displaystyle \sup_{\gamma}\displaystyle \frac{\mathrm{Ext}^{1/2}_{\gamma}(Y)}{\mathrm{Ext}^{1/2}_{\gamma}(X)}$ \\
\hline
(v) & $\| V \|_L= \displaystyle \sup_{\lambda \in \mathcal{ML}} \displaystyle \frac{d \ell_{\lambda}(V)}{\ell_{\lambda}(X)}$ & $\| V \|_T= \displaystyle \sup_{\lambda \in\mathcal{MF}} \displaystyle \frac{d \mathrm{Ext}^{1/2}_{\lambda}(V)}{\mathrm{Ext}^{1/2}_{\lambda}(X)} $ \\
\hline
(vi) & Thurston cataclysm coordinates & The homeomorphism \\
& $X \in \mathcal{T} (S) \mapsto F_\mu(X)\in\mathcal{MF}(\mu)$ & $ X\in \mathcal{T} (S) \mapsto F_h(\Phi_F(X))\in\mathcal{MF}(F)$ \\
\hline
(vii) & $\ell_\lambda(X)=i(\lambda, F_{\lambda^*}(X)) $ & $\mathrm{Ext}_{\lambda}(X)=i(\lambda, F_h(\Phi_{\lambda}(X) ))$ \\
\hline
(viii) & $d \ell_{\gamma}(\mu)=\displaystyle \frac{2}{\pi} \mathrm{Re} <\Theta_\alpha, \mu>$ & $d \mathrm{Ext}_{\lambda}(\mu)= -2 \mathrm{Re} <\Phi_\lambda, \mu>$ \\
\hline
(ix) & The horofuction boundary is & The horofuction boundary is \\
& Thurston's boundary & Gardiner-Masur's boundary \\
\hline
(x) & $\mathcal{L}_X(\lambda)=\displaystyle\frac{\ell_X(\lambda)}{L_X}$ for $\lambda\in\mathcal{ML}$ & $\mathcal{E}_X(\lambda)=\displaystyle\frac{Ext^{1/2}_X(\lambda)}{K^{1/2}_X}$ for $\lambda\in\mathcal{ML}$ \\
\hline
\end{tabular}
\end{center}
\medskip
Most of the entries in this table are well know, some of them are known but need explanation, and some of the them are new and proved in this paper.
We now make a few comments on all the entries.
\medskip
(i) Stretch maps arise from the extremal problem of finding the best Lipschitz constant of maps homotopic to the identity between two hyperbolic structures on a surface $S$, in much the same way as Teichm\"uller maps arise from the extremal problem of finding the best quasiconformal constant of maps homotopic to the identity between two complex structures on $S$.
(ii) Stretch lines are geodesics for Thurston's metric. A stretch line is determined by a pair $(\mu,F)$ where $\mu$ is a complete lamination (not necessarily measured) and $F$ a measured foliation transverse to $\mu$. Thurston proved that any two points in Teichm\"uller space can be joined by a geodesic which is a concatenation of stretch lines, but in general such a geodesic is not unique. Furthermore, there exist geodesics for Thurston's metric that are not concatenations of stretch lines. This contrasts with Teichm\"uller's theorem establishing that existence and uniqueness of geodesics joining any two distinct points, cf. \cite{T1939},
\cite{T1943} and Ahlfor's survery \cite{Ahlfors-ICM1978}.
(iii) The left hand side is Thurston's definition of Thurston's metric, the infimum is over all homeomorphisms $f$ homotopic to the identity and $L(f)$ is the Lispschitz constant of such a homeomorphism. The right hand side is the definition of the Teichm\"uller metric, the infimum is over all quasiconformal homeomorphisms $f$ homotopic to the identity and $K(f)$ is the dilatation of such a homeomorphism.
(iv) The left hand side is another expression (also due to Thurston) of Thurston's metric, and the right hand side is Kerckhoff's formula for the Teichm\"uller metric.
(v) Here, $V$ is a tangent vector to Teichm\"uller space at a point $X$. The left hand side formula is due to Thurston \cite{Thurston}; it is the infinitesimal form of Thurston's metric. The right hand side is an infinitesimal form of the Teichm\"uller metric, and it is proved below (Theorem \ref{thm:teich}).
(vi) On the left hand side, $\mu$ is a complete geodesic lamination and the range of the map, $\mathcal{MF}(\mu)$, is the subspace of $\mathcal{MF}$ of equivalence classes of measured foliations that are transverse to $\mu$.
On the right hand side, $F$ is a measured foliation, and the range of the map, $\mathcal{MF}(F)$, is the subspace of $\mathcal{MF}$ consisting of equivalence classes of measured foliations that are transverse to $F$. The left hand side map is a homeomorphism defined in Thurston's paper \cite{Thurston}. The right hand side map is a homeomorphism that arises from the fact that a pair $F_1,F_2$ of measured foliations determines a unique point $X\in\mathcal{T}(S)$ and a unique quadratic differential $\Phi$ on $X$ such that $F_1$ ad $F_2$ are measure equivalent to the vertical and horizontal foliations of $\Phi$.
Here, $F_h(\Phi_F(X))$ is the (equivalence class of the) horizontal foliation of the quadratic differential on the Riemann surface $X$ having $F$ as vertical foliation.
(vii) The left hand side is an expression of the length of a complete lamination $\lambda$ as the geometric intersection with the horocyclic foliation associated to a completion $\lambda^*$ of $\lambda$. The formula is proved in \cite{P1} for the case where $\lambda$ is complete. The case where $\lambda$ is not complete follows easily from the geometric arguments used in that proof. The right hand side formula is due to Kerckhoff, see Lemma \ref{lem:Ker} below.
(viii) In the left hand side formula, $\mu$ is a Beltrami differential. The formula was given by Gardiner \cite{Gardiner1}, and in this formula $\alpha$ is (the homotopy class of) a simple closed curve and $\Theta_\alpha$ is the Poincar\'e series of $\alpha$, that is, a quadratic differential on the hyperbolic surface $X$ defined by $$\Theta_\alpha=\sum_{B\in <A>\setminus \Gamma}B^*(\frac{dz}{z})^2$$
where $\Gamma\subset \mathrm{PSL}(2,\mathbb{R})$ is a Fuchsian group associated to $X$ acting on the upper half-plane $\mathbb{H}^2$ and $A(z)=e^{\ell_\alpha}z$ is the deck transformation corresponding to the simple closed geodesic $\alpha$.
The right hand side formula is called Gardiner's extremal length variational formula (see Lemma \ref{lem:Gardiner} below).
(ix) Walsh showed in \cite{Walsh} that the horofuction boundary of Thurston's metric is canonically identified with Thurston's boundary. Liu and Su showed in \cite{LS1} that the horofuction boundary of the Teichm\"uller metric is canonically identified with Gardiner-Masur's boundary.
(x) Here, we choose a basepoint $X_0\in \mathcal{T}(S)$ and we fix a complete geodesic lamination $\mu$ on $S$. For each $X \in \mathcal{T}(S)$, $L_X$ is the Lipschitz constant of the extremal Lipschitz map between $X_0$ and $X$ and $K_X$ is the quasiconformal dilatation of the extremal quasiconformal map between $X_0$ and $X$.
We illustrate the use of the functions $\mathcal{L}_X$ and $\mathcal{E}_X$. We use Thurston's homeomorphism $\phi_\mu: \mathcal{T} (S) \to MF(\mu)$ that maps each $ X\in \mathcal{T}(S)$ to the horocylic foliation $F_\mu(X)$. The definition of this homeomorphism is recalled in \S \ref{s:Lipschitz} below. We denote the projective class of a measured foliation $F$ by $[F]$.
By a result in \cite{P1}, a sequence $(X_k)_{k\geq 1}$ in $ \mathcal{T}(S)$ coverges to a limit $[F]\in \mathcal{PMF}$ if and only if $F_\mu(X_k)$ tends to infinity and $[F_\mu(X_k)]$ coverges to $[F]$.
We prove the following:
\begin{proposition}
If a sequence $(X_k)_{k\geq 1}$ in $ \mathcal{T}(S)$ coverges to a limit $[F]\in \mathcal{PMF}$,
then $\mathcal{L}_{X_k}(\cdot)$ converges to $[F]$ in the following sense:
\medskip
\\
$(\star)$ up to a subsequence, $\mathcal{L}_{X_k}(\cdot)$ converges to a positive multiple of $i(F,\cdot )$ uniformly on any compact subset of $ \mathcal{ML}$.
\end{proposition}
\begin{proof}
Assume that a sequence $(X_k)_{k\geq 1}$ coverges to $[F]\in \mathcal{PMF}$. Let ${L}_{X_k}=L_k$ and $F_\mu(X_k)=F_k$.
From Thurston's theory \cite{FLP}, there exists a sequence $(c_k)_{k\geq 1}$ of positive numbers such that $c_k \to 0$ and
\begin{equation}\label{equ:con1}
c_k i(F_k, \lambda) \to i(F, \lambda)
\end{equation}
for each $\lambda\in\mathcal{ML} $.
By the Fundamental Lemma in \cite{P1}, there is a uniform constant $C>0$ such that
\begin{equation}\label{equ:con2}
i(F_k, \lambda) \leq \ell_\lambda(X_k) \leq i(F_k, \lambda)+C.
\end{equation}
Note that $$c_k \ell_\lambda(X_k) \leq c_kL_k \ell_\lambda(X_0)$$
and for each $L_k$, there is a measured lamination $\lambda_k$ with $\ell_{\lambda_k}(X_0)=1$ such that $L_k= \ell_{\lambda_k}(X_k)$. As a result, we have
\begin{equation}\label{equ:con3} \frac{c_k\ell_\lambda(X_k) }{\ell_\lambda(X_0)} \leq c_kL_k\leq c_k \ell_{\lambda_k}(X_k).\end{equation}
From $(\ref{equ:con1})$, $(\ref{equ:con2})$, $(\ref{equ:con3})$, it follows that $c_kL_k$ is uniformly bounded from above and uniformly bounded below away from zero. It follows that, up to a subsequence,
$\frac{\ell_\lambda(X_k) }{L_k} $ converges to a positive multiple of $i(F, \lambda)$
for each $\lambda\in\mathcal{ML} $. Since pointed convergence of $\frac{\ell_\lambda(X_k) }{L_k} , \lambda\in\mathcal{ML} $ is equivalent to uniform convergence on compact subsets of $\mathcal{ML}$, property $(\star)$ holds.
\end{proof}
Concerning the right hand side formula in (x), a similar result was obtaind by Miyachi in \cite{Miyachi} :
\begin{proposition}
If a sequence $(X_k)_{k\geq 1}$ in $ \mathcal{T}(S)$ coverges to a limit $P$ in Gardiner-Masur's boundary, then $\mathcal{E}_{X_k}(\cdot)$ converges to some function $\mathcal{E}_P(\cdot)$ in the following sense: Up to a subsequence, $\mathcal{E}_{X_k}(\cdot)$ converges to a positive multiple of $\mathcal{E}_P(\cdot )$ uniformly on any compact subset of $ \mathcal{ML}$.
\end{proposition}
\section{Lipschitz Norm}\label{s:Lipschitz}
This section contains results of Thurston from his paper \cite{Thurston} that we will use later in this paper. We have provided proofs because at times Thurston's proofs in \cite{Thurston} are considered as sketchy.
A geodesic lamination $\mu$ on a hyperbolic surface $X$ is said to be \emph{complete} if its complementary regions are all isometric to ideal triangles. (We note that we are dealing with laminations $\mu$ that are not necessarily measured, except if specified.) Associated with $(X, \mu)$ is a measured foliation $F_{\mu}(X)$, called the horocyclic foliation, satisfying the following three properties:
\begin{enumerate}[(i)]
\item $F_{\mu}(X)$ intersects $\mu$ transversely, and in each cusp of an ideal triangle in the complement of $\mu$, the leaves of the foliation are pieces of horocycles that make right angles with the boundary of the triangle;
\item on the leaves of $\mu$, the transverse measure for $F_{\mu}(X)$ agrees with arclength;
\item there is a nonfoliated region at the centre of each ideal triangle of $S\setminus \mu$ whose boundary consists of three pieces of horocycles that are pairwise tangent (see Figure \ref{shrinkfig13bis}).
\begin{figure}[!hbp]
\centering
\psfrag{a}{\small \shortstack{horocycles\\perpendicular\\to the boundary}}
\psfrag{b}{\small \shortstack{horocycle of length 1}}
\psfrag{c}{\small \shortstack{non-foliated\\region}}
\includegraphics[width=0.5\linewidth]{shrinkfig13bis.eps}
\caption{\small {The horocyclic foliation of an ideal triangle.}}
\label{shrinkfig13bis}
\end{figure}
\end{enumerate}
We denote by $\mathcal{MF}(\mu)$ the space of measured foliations that are transverse to $\mu$. Thurston \cite{Thurston} proved the following fundamental result.
\begin{theorem}
The map $\phi_\mu: \mathcal{T} (S) \to \mathcal{MF}(\mu)$ defined by $ X \mapsto F_\mu(X)$ is a homeomorphism.
\end{theorem}
The \emph{stretch line} directed by $ \mu$ and passing through $X\in \mathcal{T} (S)$ is the curve
$$\mathbb{R} \ni t \mapsto X_t= \phi_\mu^{-1}(e^t F_\mu(X)).$$
We call a segment of a stretch line \emph{a stretch path}.
Suppose that $\mu$ is the support of a measured geodesic lamination $\lambda$.
Then, for any two points $X_s, X_t, s\leq t$ on the stretch line, their Lipschitz distance $d_L(X_s, X_t)$ is equal to $t-s$, and this distance is realized by
$$\log \frac{\ell_{\lambda} (X_t)}{\ell_{\lambda} (X_s)}.$$
We denote by $\mathcal{ML}$ the space of measured geodesic laminations on $X$ and we let $\mathcal{ML}_1= \{\lambda\in \mathcal{ML} \ | \ \ell_\lambda(X)=1 \}$. We may identify $\mathcal{ML}_1$ with $\mathcal{PL}$, the space of projective measured laminations.
Thurston \cite{Thurston} introduced a Finsler structure on $\mathcal{T}(S)$ by defining the Finsler norm of a tangent vector $V\in T_X\mathcal{T}(S)$ by the following formula :
\begin{equation}\label{eq:F}
\| V \|_L= \sup_{\lambda \in \mathcal{ML}} \frac{d \ell_{\lambda}(V)}{\ell_{\lambda}(X)} .
\end{equation}
Note that we may write
$$\| V \|_L=\sup_{\lambda \in \mathcal{ML}_1} d \ell_{\lambda}(V).$$
\begin{lemma}\label{lem:st}
Let $\Gamma_\mu(t)$ be a stretch line through $X$ with $\Gamma_\mu(0)=X$
and $\dot\Gamma_\mu(0)=V_0$.
For any measured lamination $\lambda$ supported by $\mu$ ($\lambda$ may be not unique), we have
$$\| V_0 \|_L= \frac{d \ell_{\lambda}(V_0)}{\ell_{\lambda}(X)} .$$
\end{lemma}
\begin{proof}
By multiplying the transverse measure by a constant, we may assume that $\ell_{\lambda}(X)=1$. The stretch map from $X$ to $\Gamma_\lambda(t)$ is a Lipschitz map with Lipschitz constant $e^t$, and
the hyperbolic length of $\lambda$ in $\Gamma_\lambda(t)$ is equal to $e^t$.
As a result, for any measured lamination $\gamma\in \mathcal{ML}_1$,
$$\ell_{\gamma}(\Gamma_\lambda(t))\leq e^t.$$
It follows that
$$\sup_{\gamma \in \mathcal{ML}_1} {d \ell_{\gamma}(V_0)} = d \ell_{\lambda_0}(V_0).$$
\end{proof}
\begin{theorem}[Thurston \cite{Thurston} p. 20]\label{thm:lip}
$\| \cdot \|_L $ is the infinitesimal norm of Thurston's asymmetric distance
$d_L$.
\end{theorem}
\begin{proof}
Recall that the Finsler norm $\| \cdot \|_L$ induces an (asymmetric) distance:
$$d(X,Y)= \inf_{\Gamma} \int_0^1 \| \dot\Gamma \|_L,$$
where the infimum is taken over all piecewise $C^1$ curves $\Gamma:[0,1]\to \mathcal{T}$ in Teichm\"uller space joining $X$ to $Y$.
By compactness of $\mathcal{PML}$, there exists an element $\lambda\in\mathcal{ML}$ such that the distance $d_L(X,Y)$ is attained at $\lambda$, that is,
$$d_L(X,Y)= \log \frac{\ell_{\lambda}(Y)}{\ell_{\lambda}(X)} .$$
For any piecewise $C^1$ path $\Gamma: [0, 1] \to \mathcal{T} (S)$ satisfying $\Gamma(0)=X$ and $\Gamma(1)=Y$, we have
\begin{eqnarray*}
d_L(X,Y) &=& \log \frac{\ell_{\lambda}(Y)}{\ell_{\lambda}(X)} \\
&=& \int_0^1 d \log\ell_{\lambda} ( \dot\Gamma(t) ) \\
&\leq & \int_0^1 \| \dot\Gamma(t) \|_L.
\end{eqnarray*}
Therefore, $d_L(X,Y)\leq d(X,Y). $
To prove the reverse inequality, we use a result proved by Thurston in \cite{Thurston}, namely, that any two points in Teichm\"uller space
can be connected by a finite concatenation of stretch paths, each of which stretches along some common measured geodesic lamination $\lambda$.
As a result, for any distinct $X$ and $Y$ in $\mathcal{T}(S)$, we can assume that there exists a path $\Gamma$ connecting $X$ and $Y$,
with $\Gamma$ being a concatenation $\Gamma_1 * \cdots * \Gamma_n$ of stretch paths. Up to reparametrization, we may assume that each $\Gamma_i$
is defined on $[0, 1]$. Furthermore, by a result of Thurston, we may assume that each stretch path $\Gamma_i$ is directed by a complete geodesic lamination $\mu_i$ that contains a common measured lamination $\lambda$, which is the \emph{maximally stretched lamintation} from $X$ to $Y$. This is a consequence of Theorem 8.2 in Thurston \cite{Thurston}, in which Thurston shows that there is a unique maximal ratio-maximizing chain recurrent lamination which contains all other ratio-maximizing chain recurrent lamitations.
For our purposes, we take $\lambda$ to be the maximal (with respect to inclusion) measured lamination contained in $\mu(X,Y)$.
Along each $\Gamma_i$, by Lemma \ref{lem:st}, we have
$$\sup_{\nu \in \mathcal{ML}} \frac{d \ell_{\nu}(\dot\Gamma_i)}{\ell_{\nu}} = \frac{d \ell_{\lambda}(\dot\Gamma_i)}{\ell_{\lambda}} .$$
Now we have
\begin{eqnarray*}
\int_\Gamma \| \dot\Gamma \|_L
&=& \sum_{i=1}^n \int_{\Gamma_i} \| \dot\Gamma \|_L \\
&=& \sum_{i=1}^n \int_{\Gamma_i} \frac{d \ell_{\lambda}(\dot\Gamma_i)}{\ell_{\lambda}} \\
&=& \sum_{i=1}^n \log \frac{\ell_{\lambda}(\Gamma_i(1))}{\ell_{\lambda}(\Gamma_i(0))} \\
&=& \log \frac{\ell_{\lambda}(Y)}{\ell_{\lambda}(X)} .
\end{eqnarray*}
For the last equality, note that we have $\Gamma_i(1)=\Gamma_{i+1}(0)$ for each $i$ and $\Gamma_1(0)=X, \Gamma_n(1)=Y$. It follows that $d(X,Y) \leq d_L(X,Y)$.
\end{proof}
\section{ Teichm\"uller Norm}
Extremal length is an important tool in the study of the Teichm\"uller metric. The notion is due to Ahlfors and Beurling \cite{AB1}, see also \cite{AB}. We briefly recall the definition. Given a Riemann surface $X$, a \emph{conformal metric} $\sigma$ on $X$ is a metric that is locally of the form
$\sigma(z)|dz|$ where $z$ is a local holomorphic parameter and $\sigma(z)\geq 0$ is a Borel measurable function in the local chart. We define the
$\sigma$-area of $X$ by
$$ A(\sigma)=\int_X \sigma^2(z)|dz|^2 .$$
Given a homotopy class of simple closed curves $\alpha$, its $\sigma$-length is defined by
$$L_\sigma(\alpha)=\inf_{\alpha' }\int_{\alpha'}\sigma(z)|dz|,$$
where the infimum is taken over all essential simple closed curves $\alpha' $ in the homotopy class $\alpha$. (Note that it follows from the invariance property of the expression $\sigma(z)|dz|$ that the two integrals $A(\sigma)$ and $L_\sigma(\alpha)$ are well defined, that is, they can be computed in the holomorphic local coordinates and the value obtained does not depend on the choice of the coordinates, see \cite{AB}.)
With the above notation, we define the extremal length of $\alpha$ on $X$ by
$$\mathrm{Ext}_\alpha(X)=\sup_\sigma \frac{L_\sigma^2(\alpha)}{A(\sigma)},$$
where $\sigma(z)|dz|$ ranges over all conformal metrics on $X$ satisfying
$0<A(\sigma)<\infty$.
By a result of Kerckhoff \cite{Kerckhoff}, there is a unique
continuous extension of the extremal length function to the space of measured foliations $\mathcal{MF}$,
with $\mathrm{Ext}_{a\gamma}(X)=a^2\mathrm{Ext}_{\gamma}(X)$ for $a>0$ and $\gamma$ a homotopy class of simple closed curves on $X$.
Recall that a tangent vector to Teichm\"uller space at a point $X$ can be represented by a Beltrami differential
$\mu=\mu(z)\frac{d \bar z}{d z}$ \cite{IT}. (Note that in this section we use the letter $\mu$ to denote a Beltrami differential since this is the traditional notation, although in other sections the same letter is used to denote a lamination. Hopefully, there will be no confusion.)
Using the extremal length function, we define a norm on the tangent space at $X$ by setting:
\begin{equation}\label{eq:E}
\| \mu \|_E= \sup_{\lambda \in \mathcal{MF}} \frac{d \mathrm{Ext}^{1/2}_{\lambda}(\mu)}{ \mathrm{Ext}^{1/2}_{\lambda}(X)} .
\end{equation}
The following theorem is due to Hubbard-Masur \cite{HM}; we refer to Kerckhoff \cite{Kerckhoff} for a short proof.
\begin{theorem} \label{th:HM}
For any Riemann surface $X$ and for any measured foliation $\lambda$ on $X$, there is exactly one quadratic differential, denoted by $\Phi_\lambda$, whose vertical measured foliation is measure-equivalent to $\lambda$.
\end{theorem}
\begin{lemma}[Gardiner \cite{Gardiner}] \label{lem:Gardiner}
The extremal length function $\mathrm{Ext}_{\lambda}$ is differentiable and we have the following formula, called the ``first variational formula":
\begin{equation}\label{equ:Gar}
d \mathrm{Ext}_{\lambda}(\mu)= -2 \mathrm{Re} <\Phi_\lambda, \mu>,
\end{equation}
where $<\Phi_\lambda, \mu>$ is the natural pairing
$$<\Phi_\lambda, \mu>=\iint_X \Phi_\lambda(z)\mu(z) d x d y.$$
\end{lemma}
The following observation is due to Kerckhoff \cite{Kerckhoff}:
\begin{lemma}\label{lem:Ker}
With the above notation,
\begin{equation}\label{equ:Ker}
\mathrm{Ext}_{\lambda}(X)= \iint_X |\Phi_\lambda(z)| d x d y.
\end{equation}
\end{lemma}
\begin{proof}
The proof we give here is due to Ivanov \cite{Ivanov}. We include it for completeness.
By continuity and the density of weighted simple closed curve in $\mathcal{MF}$, it suffices to prove $(\ref{equ:Ker})$ for the case where $\mu=a\gamma \in \mathcal{MF}$, where $\gamma$ is (the homotopy class of) a simple closed curve and $a >0$.
Let $\Phi$ be the one-cylinder Strebel differential on $X$ determined
by $a\gamma$. The complement of the vertical critical leaves of $\Phi$ is a cylinder foliated by circles isotopic to $\gamma$. Let us also set $\rho=|\Phi|^{1/2}|dz|$. Then $\rho$ is a flat metric on $S$, with a finite number of singular points, which are conical singularities. Measured in the flat metric $\rho$, the circumference and height of the cylinder are equal to $L_\rho(\gamma)$
and $a$ respectively.
By a theorem of Jenkins-Strebel \cite{Strebel}, the extremal length $\mathrm{Ext}_{X}(\gamma)$ of $\gamma$ is equal to
$$\mathrm{Ext}_{\gamma}(X)=\frac{L_\rho(\gamma)}{a},$$
where $\rho=|\Phi|^{1/2}|dz|$.
The area $A(\rho)=\iint_X |\Phi_\lambda(z)| d x d y$ of the cylinder is equal to $aL_\rho(\gamma)$. As a result,
$$\mathrm{Ext}_{a\gamma}(X)=a^2\mathrm{Ext}_{\gamma}(X)=a {L_\rho(\gamma)}=A(\rho).$$
\end{proof}
The next result follows from $(\ref{equ:Gar})$ and $(\ref{equ:Ker})$.
\begin{proposition}\label{pro:equality}
The norm $\| \mu \|_E$ satisfies
\begin{equation}\label{equ:norm}
\| \mu \|_E= \sup_{\lambda \in \mathcal{MF}} \frac{-\mathrm{Re} <\Phi_\lambda, \mu>}{\iint_X |\Phi_\lambda(z)| d x d y }= \sup_{\| \Phi \|=1} {\mathrm{Re} <\Phi, \mu>},
\end{equation}
where $\Phi$ varies over all holomorphic quadratic differentials $\Phi$.
\end{proposition}
Note that we have $\| \Phi \|=\iint_X |\Phi (z)| d x d y$.
\bigskip
Now we consider the Teichm\"uller metric on $\mathcal{T} (S)$.
We recall that it is defined by
$$d_T(X,Y):=\frac{1}{2} \inf_f \log K(f)$$
where $f:X \to Y$ is a
quasi-conformal map homotopic to the identity map of $S$ and
$$K(f)=\sup_{x\in X}K_x(f)\geq 1$$
is the quasi-conformal dilatation of $f$ (the $\sup$ here denotes essential supremum), with
$$K_x(f)=\frac{|f_z(x)|+|f_{\bar z}(x)|}{|f_z(x)|-|f_{\bar z}(x)|}$$
being the pointwise quasiconformal dilatation at the point $x\in X$ with local conformal
coordinate $z$.
Teichm\"uller's theorem states that given any $X,Y \in \mathcal{T}(S)$, there exists a unique
quasi-conformal map $f: X \to Y$, called the Teichm\"uller map, such that
$$d_T(X,Y)=\frac{1}{2} \log K(f).$$
The Beltrami coefficient $\mu:=\displaystyle \frac{\bar\partial f}{\partial f}$ is of the form $\mu=\displaystyle k\frac{\bar \Phi}{|\Phi|}$
for some quadratic differential $\Phi$ on $X$ and some constant $k$ with $0\leq k <1$.
In some natural coordinates given by $\Phi$ on $X$ and for some associated quadratic differential $\Phi'$ on $Y$, the Teichm\"uller map
$f$ is given by $f(x+iy)=K^{1/2}x+ i K^{-1/2}y$, where $K=K(f)=\displaystyle \frac{1+k}{1-k}$.
It is known that the Teichm\"uller metric is a Finsler metric and that between any two points in $\mathcal{T}(S)$ there is exactly one geodesic.
A geodesic ray with initial point $X$ is given by the one-parameter family of Riemann surfaces $\{X_t\}_{t\geq 0}$,
where there is a holomorphic quadratic differential $\Phi$ on $X$ and a family of
Teichm\"uller maps $f_t:X\to X_t$, with initial Beltrami differential $\mu(f_t)=\displaystyle \frac{e^{2t}-1}{e^{2t}+1}\frac{\bar \Phi}{|\Phi|}$. Here $\mu(f_t)$ is chosen such that the geodesic ray has unit speed, that is, $d_T(X_s,X_t)=t-s$ for all $s\leq t$.
We also recall the following formula due to Kerckhoff \cite{Kerckhoff}.
\begin{theorem}
Let $X, Y$ be any two points in $\mathcal {T}(S)$. Then
$$d_T(X,Y)=\frac{1}{2} \log \sup_{\lambda} {\frac{\mathrm{Ext}_{\lambda}(Y)}{\mathrm{Ext}_{\lambda}(X)}},$$
where $\lambda$ ranges over elements in $\mathcal{MF}$.
\end{theorem}
Now we can prove the following:
\begin{theorem}\label{thm:teich}
The metric induced by the norm $\| \cdot \|_E$ defined in (\ref{eq:E}) is the Teichm\"uller metric.
\end{theorem}
\begin{proof}
Denote by $d_E$ the length metric on $\mathcal {T}(S)$ induced by the norm $\| \mu \|_E$.
For any $X, Y$ in $\mathcal {T}(S)$, by Kerckhoff's formula, the Teichm\"uller distance is realized by
$$d_T(X,Y)=\frac{1}{2} \log {\frac{\mathrm{Ext}_{\lambda}(Y)}{\mathrm{Ext}_{\lambda}(X)}}$$
for some measured foliation $\lambda$. An argument similar to the one in the first part of the proof of Theorem \ref{thm:lip} shows that $d_T(X,Y)\leq d(X,Y)$.
For the converse, let $\Gamma(t)$ be a Teichm\"uller geodesic connecting $X$ and $Y$. We parametrize $\Gamma$ with unit speed, and define it on the interval $[0, T]$ with $T=d_T(X,Y)$ with $\Gamma(0)=X$ and $\Gamma(T)=Y$.
For each $t$ in $[0,T]$, the Teichm\"uller map between $X=\Gamma(0)$ and $\Gamma(t)$ has quasiconformal dilatation $e^{2t}$. It follows from the geometric definition of a quasiconformal map that for any $\lambda$ in $\mathcal{MF}$, we have
$$\mathrm{Ext}_{\lambda}(\Gamma(t)) \leq e^{2t} \mathrm{Ext}_{\lambda}(\Gamma(0)).$$
Taking square roots, we get
$$\mathrm{Ext}^{1/2}_{\lambda}(\Gamma(t)) \leq e^{t} \mathrm{Ext}^{1/2}_{\lambda}(\Gamma(0))$$
and then
\begin{eqnarray*} \lim_{t\to 0^+} \frac{\mathrm{Ext}^{1/2}_{\lambda}(\Gamma(t))-\mathrm{Ext}^{1/2}_{\lambda}(\Gamma(0))}{t \mathrm{Ext}^{1/2}_{\lambda}(\Gamma(t))}&\leq &
\lim_{t\to 0^+} \frac{e^t\mathrm{Ext}^{1/2}_{\lambda}(\Gamma(0))-\mathrm{Ext}^{1/2}_{\lambda}(\Gamma(0))}{t \mathrm{Ext}^{1/2}_{\lambda}(\Gamma(t))}\\ &=&
\lim_{t\to 0^+} \frac{e^t-1}{t}\\ &=& 1.
\end{eqnarray*}
As a result, $d \log \mathrm{Ext}^{1/2}_{\lambda} (\dot\Gamma(0)) \leq 1$ and then
$$d \log \mathrm{Ext}^{1/2}_{\lambda} (\dot\Gamma(t)) \leq 1$$ for all $t\in\mathbb{R}$.
Considering the integral of the norm $\| \cdot \|_E$ along $\Gamma(t)$, we have
$$\int_0^T \| \dot\Gamma(t) \|_E \leq T=d_T(X,Y).$$
This proves that $d(X,Y) \leq d_T(X,Y)$.
\end{proof}
Combining Proposition \ref{pro:equality} and Theorem \ref{thm:teich} , we have the following corollary.
\begin{corollary}\label{coro:teich}
The Teichm\"uller Finsler norm, denoted by $\| \mu \|_T$, is given by
$$\| \mu \|_T=\sup_{\| \Phi \|=1} \mathrm{Re} <\Phi, \mu>,$$
where $\Phi$ varies over all holomorphic quadratic differentials $\Phi$.
\end{corollary}
The above result was already known and can be obtained by using the famous Reich-Strebel inequality \cite{RS}. The result means that the Teichm\"uller norm $\| \cdot \|_T$ is dual to the $L^1$-norm
$$\| \Phi \|=\iint_X |\Phi (z)| d x d y$$
on the cotangent space.
\section{Convex embedding of measure foliation space using extremal length}
Following a usual trend, we shall call the boundary of a convex body a \emph{convex sphere}. Note that the boundary of a convex body is always homeomorphic to a sphere.
Thurston \cite{Thurston} proved that for any $X\in \mathcal{T} (S)$, the function $d \ell: \mathcal{ML}_1 \to\mathrm{T^*}_X \mathcal{T} (S)$
defined by $ \lambda\mapsto d \ell_{\lambda}$ embeds $\mathcal{ML}_1$ as a convex sphere in the cotangent space of Teichm\"uller space at $X$ containing the origin. This embedding is the dual of the boundary of the unit ball in the tangent space representing vectors of norm one for the Finsler structure associated to Thurston's metric. We prove an analogous result for the extremal length function.
\begin{theorem}
Given a point $X$ in $\mathcal{T} (S)$, let $\mathcal{PF}=\mathcal{MF}_1= \{\lambda\in \mathcal{MF} \ | \ \mathrm{Ext}_\lambda(X)=1 \}$. Then, the function $\mathcal{PF} \to \mathrm{T^*}_X \mathcal{T} (S)$ defined by
$\lambda\mapsto d \mathrm{Ext}_{\lambda}$ embeds $\mathcal{PF}$ as a convex sphere in $\mathrm{T^*}_X \mathcal{T} (S)$ containing the origin. This sphere is the boundary of the dual of the unit ball in the tangent space representing vectors of norm one for the Finsler structure associated to the Teichm\"uller metric.
\end{theorem}
\begin{proof} We fix a point $X$ in $\mathcal{T}(S)$.
By Gardiner's formula,
$$d \mathrm{Ext}_{\lambda}(\mu)= -2 \mathrm{Re} <\Phi_\lambda, \mu>$$
(see Theorem \ref{th:HM} for the notation).
Kerckhoff proved that if a sequence $\lambda_n$ in $ \mathcal{PF}$ converges to $\lambda$ then $\|\Phi_{\lambda_n}-\Phi_\lambda \| \to 0$ (see Page 34-35 of Kerckhoff's paper \cite{Kerckhoff}, where he proved the result for the case where each $\lambda_n$ is a simple closed curve. Since the subset of weighted simple closed curves is dense in $\mathcal{MF}$, the result we need follows). It follows that the map $d\mathrm{Ext}$ is continuous.
Moreover, since the pairing between quadratic differentials and harmonic Beltrami differentials is nondegenerate \cite{IT}, if $d \mathrm{Ext}_{\lambda}=d \mathrm{Ext}_{\lambda'}$, then $\Phi_{\lambda}=\Phi_{\lambda'}$.
This means that $\lambda=\lambda'$. Therefore, the map $d\mathrm{Ext}$ is injective.
Let $d\mathrm{Ext} (\mathcal{PF})$ be the image of $\mathcal{PF}$ under the map $d \mathrm{Ext}$, that is,
$$d\mathrm{Ext} (\mathcal{PF})=\{d\mathrm{Ext}_{\lambda}\in \mathrm{T^*}_X \mathcal{T} (S) \ | \ \lambda\in \mathcal{PF}\}.$$
Since $\mathcal{PF}$ is homeomorphic to a sphere, by invariance of domain, $d\mathrm{Ext} (\mathcal{PF})$ is open and $d\mathrm{Ext}$ is a homeomorphism between $\mathcal{PF}$ and $d\mathrm{Ext} (\mathcal{PF})$.
For any $\lambda\in \mathcal{PF}$, consider the Beltrami differential $$V_\lambda=-\frac{\overline{\Phi_{\lambda}}}{2|\Phi_\lambda |}.$$
It is a tangent vector in $\mathrm{T}_X \mathcal{T} (S)$. By Gardiner's formula,
\begin{eqnarray*}
d \mathrm{Ext}_\lambda(V_\lambda)&=&-2 \mathrm{Re}<\Phi_\lambda, -\frac{\overline{\Phi_{\lambda}}}{2|\Phi_\lambda |}> \\
&=&\iint_X |\Phi_\lambda| dx dy=\mathrm{Ext}_\lambda(X) =1.
\end{eqnarray*}
As a result, the derivative $d \mathrm{Ext}_{\lambda}$ in the direction $V_\lambda$ is $1$. Using H\"older's inequality,
for any $\gamma\in \mathcal{PF}$, we have
\begin{eqnarray*}
d \mathrm{Ext}_\gamma(V_\lambda)&=& \iint \Phi_\gamma \frac{\overline{\Phi_{\lambda}}}{|\Phi_\lambda |} dx dy\\
&\leq &\iint_X |\Phi_\gamma|dx dy =\mathrm{Ext}_\gamma(X)=1.
\end{eqnarray*}
Moreover, equality holds if and only if $\gamma=\lambda$.
Therefore, the derivative $d \mathrm{Ext}_{\gamma}(V_\lambda)$ of any other $\gamma\in \mathcal{PF}$ is strictly less than $1$. Denote by $C(d\mathrm{Ext} (\mathcal{PF}))$ the convex hull of $d\mathrm{Ext} (\mathcal{PF})$ in $\mathrm{T^*}_X \mathcal{T} (S)$. Thus, $V_\lambda$ defines a non-constant linear functional on the cotangent space which attains its maximal value on $C(d\mathrm{Ext} (\mathcal{PF}))$ at $d \mathrm{Ext}_{\lambda}$. As a result, $d \mathrm{Ext}_{\lambda}$ is an extreme point of the convex set $C(d\mathrm{Ext} (\mathcal{PF}))$.
We now use Thurston's argument in his proof of Theorem 5.1 of \cite{Thurston}. The set of extreme points of any convex set is a sphere of some dimension. Since the dimension of $d\mathrm{Ext} (\mathcal{PF})$ is one less than the dimension of $\mathrm{T^*}_X \mathcal{T} (S)$, the set of extremal points of $C(d\mathrm{Ext} (\mathcal{PF}))$ coincides with $d\mathrm{Ext} (\mathcal{PF})$. As a result, $d\mathrm{Ext} (\mathcal{PF})$ is a convex sphere.
To see that the convex set $C(d\mathrm{Ext} (\mathcal{PF}))$ contains the origin in its interior, note that since this convex hull has a nonempty interior, there must be at least one line through the origin of $\mathrm{T^*}_X \mathcal{T} (S)$ which intersects $d\mathrm{Ext} (\mathcal{PF})$ in at least two points.
For each of these points, there is a linear functional which attains its positive maximum value there. It follows that $0$ must separate the two points, so $0$ is a convex combination of them.
\end{proof}
\section{Comparison between the Lipschitz and Teichm\"uller norms}
From an inequality called Wolpert's inequality \cite{Wolpert}, we have $d_L \leq 2d_T$. (In fact, the inequality is contained in Sorvali's paper \cite{Sor}, and it was rediscovered by Wolpert. ) Choi and Rafi \cite{CR} proved that the two metrics are quasi-isometric in any thick part of the Teichm\"uller space, but that there are sequences $(X_n), (Y_n)$ in the thin part, with $d_L(X_n,Y_n)\to 0$.
It is interesting to compare the Lipschitz norm and the Teichm\"uller norm. In particular, we ask the following:
\begin{question} Determine a function $C(\epsilon)$ such that
$$\| V \|_L \leq \| V \|_T \leq C(\epsilon) \| V \|_L $$
for any $V\in T_X\mathcal{T} (S) $ and for all $X\in \mathcal{T}_\epsilon (S)$, the $\epsilon$-thick part of $\mathcal{T} (S)$.
\end{question}
Since $ \| V \|_L $ (respectively $ \| V \|_T$) is given by the logarithmic derivative of the hyperbolic (respectively extremal) length function of measured laminations, a further study of the relation between hyperbolic and extremal length may give the answer to the above question.
\begin{question} Consider the length-spectrum metric $d_{ls}$ on $\mathcal{T}(S)$. This can be defined as a symmetrization of Thurston's metric\footnote{Note that historically the length spectrum metric was not introduced as a symmetrization. It was first defined by Sorvali \cite{Sorvali}, before Thurston introduced his asymmetric metric.}, by the formula
\begin{eqnarray*}
d_{ls}(X,Y)&=& \max\{\log \sup_{\gamma} \frac{\ell_Y(\gamma)}{\ell_X(\gamma)}, \log \sup_{\gamma} \frac{\ell_X(\gamma)}{\ell_Y(\gamma)}\} \\
&=&\max\{d_L(X,Y), d_L(Y,X)\}. \\
\end{eqnarray*}
Is the length-spectrum metric $d_{ls}$ a Finsler metric ? If yes, give a formula for the infinitesimal norm of a vector on Teichm\"uller space with respect to this Finsler structure.
\end{question}
There are comparisons between the length-spectrum metric and the Teichm\"uller metric. It was shown by Liu and Su \cite{LS} that the divergence of $ d_{ls}$ and $d_T$ is only caused by the action of the mapping class group. In fact, they showed that $d_{ls}$ and $d_T$ are ``almost'' isometric on the moduli space and that the two metrics on the moduli space determine the same asymptotic cone.
\begin{question} One can also seek for results analogous to those presented here for the weak metric $L^ *$ dual to Thurston's metric, that is, the weak metric on Teichm\"uller space defined by
\[L^*(X,Y)=L(Y,X),\]
as well as for the length spectrum metric $d_{ls}$ on the same space.
\end{question}
|
\section{Introduction}
As alternatives to black holes, gravastars have received some attention
recently \cite{grava}, partially due to the tight connection between the
cosmological constant and a currently accelerating universe \cite{DEs},
although very strict observational constraints on the existence of such
stars may exist \cite{BN07}.
The pioneer model of gravastar was proposed by Mazur and Mottola (MM) \cite{MM01},
consisting of five layers: an internal core
$0 < R < R_1$ , described by the de Sitter universe, an intermediate thin layer of stiff fluid
$R_1 < R < R_2$ , an external region $R > R_2$ , described by the Schwarzschild solution, and two
infinitely thin shells, appearing, respectively, on the hypersurfaces $R = R_1$ and
$R = R_2$. The intermediate layer is constructed in such way that $R_1$ is inner than the de Sitter horizon, while $R_2$ is outer than the Schwarzschild horizon, eliminating the apparent horizon. Configurations with a de Sitter interior have long history which we can find, for example, in the work of Dymnikova and Galaktionov \cite{irina}.
After this work, Visser and Wiltshire \cite{VW04} pointed out that there are
two different types of stable gravastars which are stable gravastars and
``bounded excursion" gravastars. In the spherically symmetric case, the motion
of the surface of the gravastar can be written in the form \cite{VW04},
\begin{equation}
\label{1.4}
\frac{1}{2}\dot{R}^{2} + V(R) = 0,
\end{equation}
where $R$ denotes the radius of the star, and $\dot{R} \equiv dR/d\tau$, with
$\tau$ being the proper time of the surface. Depending on the properties of
the potential $V(R)$, the two kinds of gravastars are defined as follows.
{\bf Stable gravastars}: In this case, there must exist a radius $R_{0}$ such that
\begin{equation}
\label{1.5}
V\left(R_{0}\right) = 0, \;\;\; V'\left(R_{0}\right) = 0, \;\;\;
V''\left(R_{0}\right) > 0,
\end{equation}
where a prime denotes the ordinary differentiation with respect to the indicated argument.
If and only if there exists such a radius $R_{0}$ for which the above conditions are satisfied,
the model is said to be stable.
Among other things, VW found that there are many equations of state for which the gravastar
configurations are stable, while others are not \cite{VW04}. Carter studied the same
problem and found new equations of state for which the gravastar is stable \cite{Carter05},
while De Benedictis {\em et al} \cite{DeB06} and Chirenti and Rezzolla \cite{CR07}
investigated the stability of the original model
of Mazur and Mottola against axial-perturbations, and found that gravastars are stable to
these perturbations, too. Chirenti and Rezzolla also showed that their quasi-normal modes
differ from those of black holes with the same mass, and thus can be used to discern a gravastar
from a black hole.
{\bf "Bounded excursion" gravastars}: As VW noticed, there is a less stringent notion of
stability, the so-called ``bounded excursion" models, in which there exist two radii $R_{1}$
and $R_{2}$ such that
\begin{equation}
\label{1.6}
V\left(R_{1}\right) = 0, \;\;\; V'\left(R_{1}\right) \le 0, \;\;\;
V\left(R_{2}\right) = 0, \;\;\; V'\left(R_{2}\right) \ge 0,
\end{equation}
with $V(R) < 0$ for $a \in \left(R_{1}, R_{2}\right)$, where $R_{2} > R_{1}$.
Lately, we studied both types of gravastars \cite{JCAP,JCAP1}, and found that,
such configurations can indeed be constructed, although the region for the formation
of them is very small in comparison to that of black holes.
Based on the discussions about the gravastar picture some authors have proposed
alternative models \cite{Chan}. Among them, we can find a Chaplygin dark star \cite{Paramos},
a gravastar supported by non-linear electrodynamics \cite{Lobo07},
a gravastar with continuous anisotropic pressure \cite{CattoenVisser05}
and recently, Dzhunushaliev et al. worked on spherically symmetric configurations
of a phantom scalar field and they found something like a gravastar but it was unstable\cite{singleton}.
In addition, Lobo \cite{Lobo} studied two models for a dark energy fluid. One of them
describes a homogeneous energy density and the other describes an
ad-hoc monotonically decreasing energy density, although both of them are with anisotropic
pressure. In order to match an exterior Schwarzschild spacetime he
introduced a thin shell between the interior and the exterior spacetimes.
In this paper, we generalize our previous works \cite{JCAP,JCAP1} to the case where the
equation of state of the infinitely thin shell is given by $p= (1-\gamma) \sigma$ with
$\gamma$ being a constant, the interior consists of a phantom energy fluid \cite{Lobo},
while the exterior is still the Schwarzschild space.
We shall first construct three-layer dynamical models, and then show both types of
gravastars and black holes exist for various situations.
The rest of
the paper is organized as follows: In Sec. II we present the metrics of the
interior and exterior spacetimes, and write down the motion of the thin shell
in the form of equation (\ref{1.4}). In Sec. III we show the definitions of dark and
phantom energy, for the
interior fluid and for the shell. In Sec. IV we discuss the formation of
black holes from standard or phantom energy. In Sec. V we analyze the formation of
gravastar or normal star from standard or phantom energy. In Sec. VI we
study special cases where we can not have the "bounded excursion".
Finally, in Sec. VII we present our conclusions.
\section{ Dynamical Three-layer Prototype Gravastars}
The interior fluid is made of an anisotropic dark energy fluid with a metric
given by \cite{Lobo}
\begin{equation}
ds^2_{-}=-f_1 dt^2 + f_2 dr^2 + r^2 d\Omega^2,
\label{ds2-}
\end{equation}
where $d\Omega^2 \equiv d\theta^2 + \sin^2(\theta)d\phi^2$, and
\begin{eqnarray}
f_1 &=& (1+b r^2)^{\frac{1-\omega}{2}}(1+2 b r^2)^{\omega},\nonumber\\
f_2 &=& \frac{1+2 b r^2}{1 + b r^2},
\end{eqnarray}
where $\omega$ is a constant, and its physical meaning can be seen from the
following equation (\ref{prpt}). Since the mass is given by
$\bar m(r)=b r^3/[2(1+2br^2)]$ then we have that $b > 0$.
The corresponding energy density $\rho$, radial and tangential pressures $p_r$ and
$p_t$ are given, respectively, by
\begin{eqnarray}
p_r&=&\omega \rho =\left(\frac{\omega
b}{8\pi}\right)\left(\frac{3+2b r^2}{(1+2b r^2)^2}\right), \nonumber \\
p_t&=& -\left(\frac{b}{8\pi}\right)\left(\frac{\omega(3+2b
r^2)}{(1+2b r^2)^2}\right)
+ \frac{b^2r^2}{32\pi\left[(1+2b r^2)^3(1+b r^2)\right]}\times \nonumber \\
&&\Big\{(1+\omega)(3+2b r^2)\left[(1+3\omega)+2br^2(1+\omega)\right] \nonumber \\
&&-8\omega(5+2br^2)(1+br^2)\Big\}.
\label{prpt}
\end{eqnarray}
The exterior spacetime is given by the Schwarzschild metric
\begin{equation}
ds^2_{+}= - f dv^2 + f^{-1} d{\bf r}^2 + {\bf r}^2 d\Omega^2,
\label{ds2+}
\end{equation}
where $f=1 - {2m}/{\bf r}-({\bf r}/L_e)^2$ and $L_e=\sqrt{3/\Lambda_e}$.
The metric of the hypersurface on the shell is given by
\begin{equation}
ds^2_{\Sigma}= -d\tau^2 + R^2(\tau) d\Omega^2.
\label{ds2Sigma}
\end{equation}
Since $ds^2_{-} = ds^2_{+} = ds^2_{\Sigma}$, we find that $r_{\Sigma}={\bf r}_{\Sigma}=R$,
and
\begin{eqnarray}
\label{dott2}
f_1\dot t^2 - f_2 \dot R^2 &=& 1,\\
\label{dotv2}
f\dot v^2 - \frac{\dot R^2}{f} &=& 1,
\end{eqnarray}
where the dot denotes the ordinary differentiation with respect to the proper time.
On the other hand, the interior and exterior normal vectors to the thin shell are given by
\begin{eqnarray}
\label{nalpha-}
n^{-}_{\alpha} &=& (-\dot R, \dot t, 0 , 0 ),\nonumber\\
n^{+}_{\alpha} &=& (-\dot R, \dot v, 0 , 0 ).
\end{eqnarray}
Then, the interior and exterior extrinsic curvature are given by
\begin{eqnarray}
K^{i}_{\tau\tau}&=&\frac{1}{2} (1+b R^2)^{-\omega/2} \dot t \left\{ \left[ 4 (1+b R^2
)^{\omega/2} b R^2 \dot R^2+2 (1+b R^2)^{\omega/2} \dot R^2- \right. \right. \nonumber \\
& &\left. \left. (1+2 b R^2)^\omega \sqrt{1+b R^2} b R^2 \dot t^2-(1+2 b R^2)^
\omega \sqrt{1+b R^2} \dot t^2\right] (2 b R^2 \omega+2 b R^2+3 \omega+1)- \right. \nonumber\\
& &\left. 2 (1+b R^ 2)^{\omega/2} (1+2 b R^2) \dot R^2 \right\} (1+2 b R^2
)^{-2} (1+b R^2)^{-1} b R+ \dot R \ddot t- \ddot R \dot t,\\
\label{Ktautau-}
\end{eqnarray}
\begin{equation}
K^{i}_{\theta\theta} = \frac{\dot t(1+b R^2) R}{1 + 2 b R^2},
\label{Kthetatheta-}
\end{equation}
\begin{equation}
K^{i}_{\phi\phi} = K^{-}_{\theta\theta}\sin^2(\theta),
\label{Kphiphi-}
\end{equation}
\begin{eqnarray}
K^{e}_{\tau\tau}&=&\dot v [(2 L_e^2 m \dot v+L_e^2 R \dot R-L_e^2 R \dot v+R^
3 \dot v) (2 L_e^2 m \dot v-L_e^2 R \dot R-L_e^2 R \dot v+R^3 \dot v)- \nonumber \\
& &2 L_e^4 R^2 \dot R^2] ((2 m-R) L_e^2+R^ 3)^{-1}
(L_e^2 m-R^3) L_e^{-4} R^{-3}+\dot R \ddot v- \ddot R \dot v
\label{Ktautau+}
\end{eqnarray}
\begin{equation}
K^{e}_{\theta\theta}= -\dot v((2 m-R) L_e^2+R^3) L_e^{-2}
\label{Kthetatheta+}
\end{equation}
\begin{equation}
K^{e}_{\phi\phi}=K^{e}_{\theta\theta}\sin^2(\theta).
\label{Kphiphi+}
\end{equation}
Since \cite{Lake}
\begin{equation}
[K_{\theta\theta}]= K^{e}_{\theta\theta}-K^{i}_{\theta\theta} = - M,
\label{M}
\end{equation}
where $M$ is the mass of the shell, we find that
\begin{equation}
M=\dot v (2 m-R)+\frac{\dot t(1+b R^2) R}{1 + 2 b R^2}.
\label{M1}
\end{equation}
Then, substituting equations (\ref{dott2}) and (\ref{dotv2}) into (\ref{M1})
we get
\begin{equation}
M=-R\left(1-\frac{2m}{R} -\left(\frac{R}{L_e}\right)^2 + \dot R^2 \right)^{1/2} +
R\frac{ \left[ 1 + b R^2 + \dot R^2 (1 + 2 b R^2) \right]^{1/2}}
{(1+b R^2)^{-(\omega+1)/4}(1+2b R^2)^{(\omega+2)/2}}.
\label{M2}
\end{equation}
In order to keep the ideas of MM as much as possible, we consider the thin
shell as consisting
of a fluid with the equation of state, $p=(1-\gamma)\sigma$, where $\sigma$ and $p$ denote,
respectively, the surface energy density and pressure of the shell and $\gamma$ is a constant.
Then, the equation of motion of the shell is given by \cite{Lake}
\begin{equation}
\dot M + 8\pi R \dot R p = 4 \pi R^2 [T_{\alpha\beta}u^{\alpha}n^{\beta}]=
4\pi R^2 \left(T^e_{\alpha\beta}u_e^{\alpha}n_e^{\beta}-T^i_{\alpha\beta}u_i^{\alpha}n_i^{\beta} \right),
\label{dotM}
\end{equation}
where $u^{\alpha}$ is the four-velocity. Since the interior fluid is made
of an anisotropic fluid and the exterior is vacuum, we get
\begin{equation}
\dot M + 8\pi R \dot R (1-\gamma)\sigma = 0.
\label{dotM1}
\end{equation}
Recall that $\sigma = M/(4\pi R^2)$, we find that equation (\ref{dotM1}) has the solution
\begin{equation}
M=k R^{2(\gamma-1)},
\label{Mk}
\end{equation}
where $k$ is an integration constant. Substituting equation (\ref{Mk}) into equation (\ref{M2}),
and rescaling $m, \; b$ and $R$ as,
\begin{eqnarray}
m &\rightarrow& mk^{-\frac{1}{2\gamma-3}},\nonumber\\
b &\rightarrow& b k^{\frac{2}{2\gamma-3}},\nonumber\\
R &\rightarrow& Rk^{-\frac{1}{2\gamma-3}},
\end{eqnarray}
we find that it can be written in the form of equation (\ref{1.4}), and
\begin{eqnarray}
V(R,m,L_e,\omega,b,\gamma)&=& -\frac{1}{2L_e^2 R^2 b_2 \left( b_2^{\omega+1} b_1^{-\frac{1}{2} \omega}-b_1^{\frac{1}{2}} \right)^2} (b_2^{2 \omega+3} R^{4 \gamma-4} b_1^{-\omega} L_e^2 \nonumber \\
& &-2 b_1^{-\frac{1}{2} \omega+\frac{1}{2}} R^{2 \gamma-2} b_2^{\omega+1} L_e (-b_2 (b_2^{\omega+2} b_1^{-\frac{1}{2} \omega} R^6 L_e^2-2 b_2^{\omega+2} b_1^{-\frac{1}{2} \omega} m L_e^2 R^5- \nonumber \\
& &b_2^{\omega+2} b_1^{-\frac{1}{2} \omega} R^8-b_2^{\omega+1} b_1^{-\frac{1}{2} \omega+1} R^6 L_e^2-R^6 L_e^2 b_2 b_1^{\frac{1}{2}}+2 m L_e^2 b_2 R^5 b_1^{\frac{1}{2}}+ \nonumber \\
& &R^8 b_2 b_1^{\frac{1}{2}}+R^6 L_e^2 b_1^{\frac{3}{2}}-b_2^{\omega+2} L_e^2 b_1^{-\frac{1}{2} \omega} R^{4 \gamma})/(b_1^{1/2} R^4))^{\frac{1}{2}}+b_1^{\frac{3}{2}-\frac{1}{2} \omega} b_2^{\omega+1} R^2 L_e^2- \nonumber \\
& &b_1^2 R^2 L_e^2+b_2^{\omega+2} b_1^{-\frac{1}{2} \omega+\frac{1}{2}} L_e^2 R^{4 \gamma-4}-b_2^{2 \omega+3} L_e^2 b_1^{-\omega} R^2+ b_2^{\omega+2} b_1^{-\frac{1}{2} \omega+\frac{1}{2}} R^2 L_e^2+ \nonumber \\
& &2 b_2^{2 \omega+3} b_1^{-\omega} R m L_e^2-2 b_2^{\omega+2} b_1^{-\frac{1}{2} \omega+\frac{1}{2}} m L_e^2 R+b_2^{2 \omega+3} b_1^{-\omega} R^4-b_2^{\omega+2} b_1^{-\frac{1}{2} \omega+\frac{1}{2}} R^4) \nonumber \\
\label{VR}
\end{eqnarray}
where
\begin{equation}
\label{b1}
b_1 \equiv 1+b R^2,\;\;\;
b_2 \equiv 1+2 b R^2.
\end{equation}
The exterior horizons are given by \cite{Shankaranarayanan}
\begin{equation}
r_{bh}= \frac{2m}{\sqrt{3 y}} \cos \left( \frac{\pi+\psi}{3} \right),
\label{rbh}
\end{equation}
\begin{equation}
r_c= \frac{2m}{\sqrt{3 y}} \cos \left( \frac{\pi-\psi}{3} \right),
\label{rc}
\end{equation}
where $y=(m/L_e)^2$, $\psi= \arccos \left( 3\sqrt{3 y} \right)$, $r_{bh}$ denotes the black
hole horizon and $r_c$ denotes the cosmological horizon.
Note that if $y > 1/27$
the quantity $3\sqrt{3 y}$ is greater than 1, giving an imaginary angle $\psi$.
Thus, the
horizons $r_{bh}$ and $r_c$ are imaginary and the spacetime becomes free of
horizons.
To guarantee
that initially the spacetime does not have any kind of horizons, cosmological or event,
we must restrict $R_{0}$ to the range,
\begin{equation}
\label{2.2b}
r_{bh} < R_{0} < r_h \; or \; r_c,
\end{equation}
where $R_0$ is the initial collapse radius.
Clearly, for any given constants $m$, $\omega$, $b$ and $\gamma$, equation (\ref{VR}) uniquely
determines the collapse of the prototype gravastar. Depending on the initial value $R_{0}$,
the collapse can form either a black hole, a gravastar, a Minkowski, or a spacetime filled with
phantom fluid. In the last case, the thin shell
first collapses to a finite non-zero minimal radius and then expands to infinity. To guarantee
that initially the spacetime does not have any kind of horizons, cosmological or event,
we must restrict $R_{0}$ to the range,
\begin{equation}
\label{2.2c}
R_{0} > 2 m,
\end{equation}
where $R_0$ is the initial collapse radius. When $m = 0= b$, the thin shell disappears,
and the whole spacetime is Minkowski. So, in the following we shall not consider this case.
Since the potential (\ref{VR}) is so complicated, it is too difficult to study it
analytically. Instead, in the following we shall study it numerically. Before doing so, we
shall show the classifications of matter, dark energy, and phantom energy for
anisotropic fluids.
\section{Classifications of Matter, Dark Energy, and Phantom Energy for Anisotropic Fluids}
Recently \cite{Chan08}, the classification of matter, dark and phantom energy
for an anisotropic fluid was given in terms of the energy conditions. Such a classification
is necessary for systems where anisotropy is important, and the pressure components
may play very important roles and can have quite different contributions.
In this paper, we will use this classification to study the collapse of the
dynamical prototype gravastars, constructed in the last section.
In particular, we define dark
energy as a fluid which violates the strong energy
condition
(SEC). From the Raychaudhuri equation, we can see that such defined
dark energy always exerts divergent forces on time-like or null geodesics.
On the other hand, we define phantom energy as a fluid that violates at least
one of the null energy conditions (NEC's). We shall further distinguish phantom
energy that satisfies the SEC
from that which does not satisfy the SEC. We call
the former attractive
phantom energy, and the latter
repulsive phantom energy.
Such a classification is summarized in Table I.
For the sake of completeness, in Table II we apply it to the matter field
located on the thin shell, while in Table III we combine all the results of Tables I
and II, and present all the possibilities.
\begin{table}
\caption{\label{tab:table1} This table summarizes the classification of
the interior matter field, based on the energy conditions \cite{HE73}, where
we assume that $\rho \ge 0$.}
\begin{ruledtabular}
\begin{tabular}{cccc}
Matter & Condition 1 & Condition 2 & Condition 3 \\
\hline
Normal Matter & $\rho+p_r+2p_t\ge 0$ & $\rho+p_r\ge 0$ & $\rho+p_t\ge 0$ \\
\hline
Dark Energy & $\rho+p_r+2p_t < 0$ & $\rho+p_r\ge 0$ & $\rho+p_t\ge 0$ \\
\hline
& & $\rho+p_r < 0$ & $\rho+p_t\ge 0$ \\
Repulsive Phantom Energy & $\rho+p_r+2p_t < 0$ & $\rho+p_r\ge 0$ & $\rho+p_t < 0$ \\
& & $\rho+p_r < 0$ & $\rho+p_t < 0$ \\
\hline
Attractive Phantom Energy & $\rho+p_r+2p_t\ge 0$ & $\rho+p_r\ge 0$ & $\rho+p_t < 0$ \\
& & $\rho+p_r < 0$ & $\rho+p_t\ge 0$ \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{table}
\caption{\label{tab:table2} This table summarizes the classification of matter
on the thin shell, based on the energy conditions \cite{HE73}. The last column indicates
the particular values of the parameter $\gamma$, where we assume that $\rho \ge 0$.}
\begin{ruledtabular}
\begin{tabular}{cccc}
Matter & Condition 1 & Condition 2 & $\gamma$ \\
\hline
Normal Matter & $\sigma+2p\ge 0$ & $\sigma+p\ge 0$ & -1 or 0 \\
Dark Energy & $\sigma+2p < 0$ & $\sigma+p\ge 0$ & 7/4 \\
Repulsive Phantom Energy & $\sigma+2p < 0$ & $\sigma+p < 0$ & 3 \\
\end{tabular}
\end{ruledtabular}
\end{table}
In order to consider the equations (\ref{ds2-}) and (\ref{prpt}) for describing dark energy
stars we must analyze carefully the ranges of the parameter $\omega$ that in
fact furnish the expected fluids. It can be shown that the
condition $\rho+p_r>0$ is violated for $\omega<-1$ and fulfilled for $\omega>-1$,
for any values of $R$ and $b$.
The conditions $\rho+p_t>0$ and $\rho+p_r+2p_t>0$ are satisfied for $\omega<-1$
and $-1/3<\omega<0$, for any values of $R$ and $b$.
For the other intervals of
$\omega$ the
energy conditions depend
on very complicated relations of $R$ and $b$. See reference \cite{Chan08}.
This provides an explicit example, in which the definition of dark energy must
be dealed with great care. Another case was provided in a previous work \cite{Chan08}.
Taking several values of $\omega$ in the intervals $-1<\omega<-1/3$ and
$\omega>0$, we could not found any case where the interior dark energy exist.
In order to fulfill the energy condition $\sigma+2p\ge0$ of the shell
and assuming that
$p=(1-\gamma)\sigma$ we must have $\gamma \le 3/2$. On the other hand, in order
to satisfy the condition $\sigma+p\ge 0$, we obtain $\gamma \le 2$.
Hereinafter, we will use only some particular values of the parameter
$\gamma$ which are analyzed in this work. See Table II.
\section{Structures Formed}
Here we can find many types of systems, depending on the combination of the
constitution matter of the shell and core. Among them, there are formation of
black holes, stable and "bounded excursion" gravastars, as it has already
shown in our previous works \cite{JCAP}-\cite{JCAP4},
and even a naked singularity constituted exclusively of dark energy.
All of them are listed in the table III.
As can be seen in the figures \ref{fig1}, \ref{fig3} and \ref{fig5}, depending
on the value of the cosmological constant, we can see that $V(R) = 0$ now can i
have one, two or three real roots. Then, we
have, say, $R_{i}$, where $R_{i+1} > R_{i}$. For $L_e=L_1$ (corresponding to
$\Lambda=0$) If we choose $R_{0} > R_{3}$ none structure is allowed in this
region because the potential is greater than the zero. However, if we choose
$R_{2} < R_{0} < R_{3}$, the collapse will bounce back and forth between
$R = R_{1}$ and $R = R_{2}$. This is exactly the so-called "bounded excursion"
model mentioned in \cite{VW04}, and studied in some details in
\cite{JCAP}-\cite{JCAP4}. Of
course, in a realistic situation, the star will emit both gravitational waves
and particles, and the potential will be self-adjusted to produce a minimum at
$R = R_{static}$ where $V\left(R=R_{static}\right) = 0 = V'\left(R=R_{static}\right)$
whereby a gravastar is finally formed \cite{VW04,JCAP,JCAP1,JCAP2}.
For $R_{0} < R_{1}$ a black hole is formed in the end of the collapse of the shell.
The scenario above can significantly be changed if we consider $\Lambda>0$.
In this case for $L_e>L_c$, we also have bounded excursion gravastars if
$R_{2} < R_{0} < R_{3}$. However, for $R_{0} < R_{1}$ the final structure
can be now a black hole or a naked singularity since the presence of the
cosmological constant above a certain limit ($L_e^*$) eliminates the
event horizon (its radius becomes imaginary), as can be seen in the tables IV,
V and VI. This is the first
evidence of a naked singularity formation from a gravastar model. Moreover
for $L_e=L_c$, then $R_{2} = R_{3}$, a stable gravastar is formed if
$R_{0} = R_{2}$, while for $L_e<L_c$ there is only one real root. Note that
for any value of $L_e>L_e^*$, a naked singularity is formed for small
initial radius of the shell.
This is already present in the de Sitter-Schwarzschild solution
\cite{Shankaranarayanan}, since is the exterior cosmological constant which allows
to relax the inevitability of the horizon formation (differently from the
Schwarzschild solution). The news here are that we have found a source (the shell
with an interior anisotropic dark energy fluid) which can be matched with the de
Sitter-Schwarzschild vacuum
spacetime and, depending on the values of the exterior cosmological
constant and the total mass, can represent a naked singularity.
Thus, solving equation (\ref{M2}) for $\dot R(\tau)$ we can integrate
$\dot R(\tau)$ and obtain $R(\tau)$, which are
shown in the figures \ref{fig2}, \ref{fig4} and \ref{fig6} for the case G.
\begin{table}
\caption{\label{tab:table3}This table summarizes all possible kind of energy
of the interior fluid and of the shell and compares the formed structures
in the two gravastar models ($\Lambda_e=0$, $\Lambda_e>0$). The letters SG, UG, BEG, BH, NS and N
denote stable gravastar, unstable gravastar, bounded excursion gravastar,
black hole, naked singularity and none, respectively.}
\begin{ruledtabular}
\begin{tabular}{cccccc}
Case & Interior Energy & Shell Energy & Figures & Structures ($\Lambda_e=0$) & Structures ($\Lambda_e>0$)\\
\hline
A & Standard & Standard & & SG & SG/UG/BEG \\
B & Standard & Dark & & BH & SG/UG/BEG \\
C & Standard & Repulsive Phantom & & BH & BH \\
D & Dark & Standard & & N & N \\
E & Dark & Dark & & N & N \\
F & Dark & Repulsive Phantom & & N & N \\
G & Repulsive Phantom & Standard & \ref{fig1}, \ref{fig3}, \ref{fig5} & SG/BEG & BH/SG/UG/BEG/NS \\
H & Repulsive Phantom & Dark & & BH & BH/SG/UG/BEG \\
I & Repulsive Phantom & Repulsive Phantom & & BH & BH \\
J & Attractive Phantom & Standard & & SG/BEG & BH/SG/BEG \\
K & Attractive Phantom & Dark & & BH & BH \\
L & Attractive Phantom & Repulsive Phantom & & BH & BH \\
\end{tabular}
\end{ruledtabular}
\end{table}
\section{Conclusions}
In this paper, we have studied the problem of the stability of gravastars by
constructing dynamical three-layer models of VW \cite{VW04},
which consists of an internal anisotropic dark energy fluid, a dynamical
infinitely thin shell of perfect fluid with the equation of state
$p = (1-\gamma)\sigma$, and an external de Sitter-Schwarzschild spacetime.
We have shown explicitly that the final output can be a black
hole, a "bounded excursion" stable gravastar depending on the total mass
$m$ of the system, the cosmological constant $L_e$, the parameter $\omega$,
the constant $b$, the parameter $\gamma$ and
the initial position $R_{0}$ of the dynamical shell. All these possibilities
have non-zero measurements in the phase space of $m$, $L_e$, $b$, $\omega$,
$\gamma$ and $R_{0}$. All the results can be summarized in Table III.
An interesting result that we can deduce from Table III is that we can have
black hole and stable gravastar formation even with an interior and a shell
constituted of dark and repulsive dark energy (cases H and I).
Still more interesting
is the case G, represented by figures \ref{fig1}, \ref{fig3} and \ref{fig5}
where for small radius of the shell we
have no formation of a black hole (for $\Lambda=0$) and a naked singularity
(for $\Lambda > 0$). This is the first time in the literature that a naked
singularity emerges from a gravastar model. Besides, the figures \ref{fig2},
\ref{fig4} and \ref{fig6} give us examples of the dynamical evolution of a
gravastar to a naked singularity.
Finally, the opposite final fates, because of the absence or the presence
of a positive cosmological constant,
reinforces the hypothesis proposed in \cite{nakedpressure}
that can exist a connection between naked singularities and some kind of
weakness of the gravitational field, compared to that associated to black holes.
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=fig22grav3variandoLext.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=ECw1v5b0v01.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{The potential $V(R)$ and the energy conditions EC1$\equiv \rho+p_r+2p_t$,
EC2$\equiv \rho+p_r$ and EC3$\equiv \rho+p_t$, for $\gamma=-1$,
$\omega=1.5$, $b=0.01$ and $m_c=0.2097639045$. {\bf Case G}}
\label{fig1}
\end{figure}
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=dotRvstaum0v2097639045.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=Rvstaum0v2097639045.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{These figures represent the dynamical evolution of the shell to a
naked singularity for the potential given by the figure \ref{fig1},
assuming $R_0=R(0)=0.7$.}
\label{fig2}
\end{figure}
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=fig23grav3variandoLext.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=ECw1v5b0v01.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{The potential $V(R)$ and the energy conditions EC1$\equiv \rho+p_r+2p_t$,
EC2$\equiv \rho+p_r$ and EC3$\equiv \rho+p_t$, for $\gamma=0$,
$\omega=1.5$, $b=0.01$ and $m_c=0.3775$. {\bf Case G}}
\label{fig3}
\end{figure}
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=dotRvstaum0v3775.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=Rvstaum0v3775.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{These figures represent the dynamical evolution of the shell to a
naked singularity for the potential given by the figure \ref{fig3},
assuming $R_0=R(0)=1$.}
\label{fig4}
\end{figure}
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=fig26grav3variandoLext.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=ECw3v0b0v01.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{The potential $V(R)$ and the energy conditions EC1$\equiv \rho+p_r+2p_t$,
EC2$\equiv \rho+p_r$ and EC3$\equiv \rho+p_t$, for $\gamma=-1$,
$\omega=3$, $b=0.01$ and $m_c=0.2552945103$. {\bf Case G}}
\label{fig5}
\end{figure}
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=dotRvstaum0v2552945103.eps,width=3.3truein,height=3.0truein}\hskip
.25in \psfig{figure=Rvstaum0v2552945103.eps,width=3.3truein,height=3.0truein}
\hskip .5in} \caption{These figures represent the dynamical evolution of shell to a
naked singularity for the potential given by the figure \ref{fig5},
assuming $R_0=R(0)=0.9$.}
\label{fig6}
\end{figure}
\begin{table}
\caption{\label{tab:tablea} This table show the calculated horizons using
the equations \ref{rbh} and \ref{rc}. The symbol $i=\sqrt{-1}$ denotes the imaginary constant. See figure \ref{fig1}.}
\begin{ruledtabular}
\begin{tabular}{cccc}
$m$ & $L_e$ & Cosmological Horizon & Black Hole Horizon \\
\hline
0.2097639045 & 0.8 & 0.4795669330+0.2235024144 $i$ & 0.4795669330-0.2235024144 $i$ \\
& $L_e^*$ & & \\
& 1.2 & 0.8577020327 & 0.5136246691 \\
& 1.3104314 & 0.9974942529 & 0.4866382631 \\
& 1.5 & 1.213250267 & 0.4638970952 \\
& 10 & 9.783239233 & 0.4202701210 \\
& $\infty$ & $\infty$ & 2m=0.4195278090 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{table}
\caption{\label{tab:tableb} This table show the calculated horizons using
the equations \ref{rbh} and \ref{rc}. The symbol $i=\sqrt{-1}$ denotes the imaginary constant. See figure \ref{fig3}.}
\begin{ruledtabular}
\begin{tabular}{cccc}
$m$ & $L_e$ & Cosmological Horizon & Black Hole Horizon \\
\hline
0.3775 & 1.5 & 0.8943821768+0.3869863502 $i$ & 0.8943821767-0.3869863502 $i$ \\
& $L_e^*$ & & \\
& 1.97 & 1.1976796420 & 1.0759952780 \\
& 2.0053430150 & 1.2948106450 & 1.0151227670 \\
& 3 & 2.5081539280 & 0.8151917235 \\
& 20 & 19.61123828 & 0.7560805420 \\
& $\infty$ & $\infty$ & 2m=0.7550 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{table}
\caption{\label{tab:tablec} This table show the calculated horizons using
the equations \ref{rbh} and \ref{rc}. The symbol $i=\sqrt{-1}$ denotes the imaginary constant. See figure \ref{fig5}.}
\begin{ruledtabular}
\begin{tabular}{cccc}
$m$ & $L_e$ & Cosmological Horizon & Black Hole Horizon \\
\hline
0.2552945103 & 1.0 & 0.5973628039+0.2655691220 $i$ & 0.5973628039 - 0.2655691220 $i$ \\
& $L_e^*$ & & \\
& 1.36 & 0.8837041274 & 0.6823811248 \\
& 1.430886 & 1.002124260 & 0.6365867478 \\
& 1.6 & 1.220128096 & 0.5913790740 \\
& 12 & 11.73606139 & 0.5115184600 \\
& $\infty$ & $\infty$ & 2m=0.5105890206 \\
\end{tabular}
\end{ruledtabular}
\end{table}
\begin{acknowledgments}
The financial assistance from FAPERJ/UERJ (MFAdaS) are gratefully acknowledged.
The authors (RC, MFAdaS, JFVR) acknowledges the financial support from FAPERJ (no. E-26/171.754/2000,
E-26/171.533/2002, E-26/170.951/2006, E-26/110.432/2009 and E26/111.714/2010). The authors (RC,
MFAdaS and JFVdR) also acknowledge the financial support from Conselho Nacional de Desenvolvimento Cient\'ifico e
Tecnol\'ogico - CNPq - Brazil (no. 450572/2009-9, 301973/2009-1 and 477268/2010-2). The author (MFAdaS)
also acknowledges the financial support from Financiadora de Estudos e Projetos - FINEP - Brazil
(Ref. 2399/03).
\end{acknowledgments}
|
\section{Introduction}
\label{sec:intro}
The term ``cubature'' indicates the numerical computation of a multiple integral. This is an important topic in many different disciplines, with a correspondingly large body of literature. A description of the different kinds of cubature rules that exist, as well as of the mathematics used to derive them, is given in the classical book of \citet{Stroud1971}, with more updated information to be found, among others, in \citep{Cools1997} and in chapter~6 of~\citep{KrommerUeberhuber1998}. \Citet{Stroud1971} also presents a compilation of known (at the time) cubature rules, while newer rules are catalogued in \citep{CoolsRabinowitz1993,Cools1999} and online at the Encyclopedia of Cubature Formulas~\citep{Cools2003}.
A commonly used method to derive specific cubature rules is based on moment equations and invariant theory \citep[pp.~170--182]{KrommerUeberhuber1998}. This method, which will be used in the present paper, exploits symmetries and invariant theory to set up a non-linear system of equations, whose unknowns are the positions and weights of the integration points. The use of invariants, together with appropriate analytical computations, can lead to a significant simplification of the system of equations, which however in most cases still has to be solved numerically.
Although appropriate iterative numerical methods have been successfully used to obtain \emph{individual} numerical solutions to the aforementioned system of equations, obtaining a solution in this way provides no information on its uniqueness. Conversely, inability to obtain a solution does not prove its inexistence (though it is a strong indication, when sufficiently robust numerical methods are employed). It is thus interesting and useful to be able to compute analytically all the solutions for a cubature rule.
In this paper we focus on fully symmetric cubature rules on the triangle and provide results of the analytical computation for cubature rules of moderate degree, extending significantly the analytical results given by \citet{LynessJespersen1975}. The theory of symmetric polynomials~\citep{Macdonald1998} is used in the generation of the non-linear system of equations, resulting in a straightforward way to formulate a system which is simpler than those resulting from the use of polar coordinates~\citep{LynessJespersen1975,BerntsenEspelid1990} or Cartesian coordinates~\citep{WandzuraXiao2003}. Additionally, symmetric polynomials are also used to significantly simplify the computation and presentation of the analytical solution.
\section{Symmetric polynomials}
Although the formulation presented here is actually based on invariant theory, the relevant theory is not used directly but is ``implied'' by using the theory of symmetric polynomials~\citep{Macdonald1998}. As we will see in the following, the use of symmetric polynomials provides a concise formulation of the non-linear system of equations, while also leading to simpler computation and presentation of the solution.
A \emph{symmetric polynomial} is a multivariate polynomial in $n$ variables, say $x_1,x_2,\ldots,x_n$, which is invariant under any permutation of its variables. We define the \emph{elementary symmetric polynomials} $\tilde{x}_k$ as the sums of all products of $k$ distinct variables $x_i$, with negative sign when $k$ is odd, that is
\begin{equation}
\label{eq:elemsym}
\tilde{x}_k = (-1)^k \sum_{i_1<i_2<\ldots<i_k} x_{i_1} x_{i_2} \cdots x_{i_k}
\end{equation}
with $\tilde{x}_0 = 1$. The alternating sign $(-1)^k$ in equation~(\ref{eq:elemsym}), which does not appear in the usual definition of the elementary symmetric polynomials, is introduced here as it leads to simpler expressions. While elementary symmetric polynomials are usually denoted using a letter (e.g.\ $\Pi_k$, $s_k$ or $e_k$) which is different from the variable name, we use here the superimposed tilde over the variable name since we will be dealing with elementary symmetric polynomials of different sets of variables.
The \emph{fundamental theorem of symmetric polynomials} states that any symmetric polynomial in the variables $x_i$ can be expressed as a polynomial in the elementary symmetric polynomials $\tilde{x}_k$.
Equation~(\ref{eq:elemsym}) allows computing the elementary symmetric polynomials $\tilde{x}_k$ in terms of the $n$ variables $x_i$. Conversely, the values $x_i$ can be calculated from $\tilde{x}_k$ as the solutions for $x$ of the polynomial equation
\begin{equation}
\sum_{j=0}^{n} \tilde{x}_{n-j} x^j = 0
\end{equation}
\section{Formulating the system of equations}
\label{sec:formulating}
\subsection{Moment equations}
\label{sec:momeqs}
Our objective is to derive a cubature formula (or \emph{rule}) for the approximate evaluation of the integral of a function $f$ over the area $A$ of a triangle
\begin{equation}
\label{eq:gencub}
\bar{I} = \sum_{i=1}^{n_p} \bar{w}_i f^{(i)}
\approx \frac{1}{A} \int_A f \,\mathrm{d} A
\end{equation}
where $f^{(i)}$ is the value of $f$ at point $i$, $\bar{w}_i$ is the corresponding weight and $n_p$ is the number of points used in the cubature. We only consider rules of (polynomial) degree $d$, that is rules where equation~(\ref{eq:gencub}) is exact for all polynomials of degree less or equal to $d$, while it is not exact for at least one polynomial of degree $d+1$.
A polynomial of degree $d$ on the triangle can be written as a linear combination of terms $L_1^i L_2^j L_3^{d-i-j}$, where $L_1$, $L_2$ and $L_3$ are the areal coordinates. The cubature rule can therefore be determined by requiring that equation~(\ref{eq:gencub}) is exact for each of these terms. The resulting equations are known as the \emph{moment equations}. The number $\bar{n}_e$ of different terms $L_1^i L_2^j L_3^{d-i-j}$, which is the number of equations to be solved, is
\begin{equation}
\label{eq:ne_bar}
\bar{n}_e = (d+1)(d+2)/2
\end{equation}
We only consider \emph{fully symmetric} rules where, if a point with areal coordinates ($L_1$,$L_2$,$L_3$) is used in the cubature, then all points resulting from the permutation of the areal coordinates are also used, with the same weight.
Integration points in a fully symmetric rule can thus belong to one of three different types of point sets, or \emph{orbits}, depending on the number of areal coordinates which are equal. If all areal coordinates are equal, we get a single ``type-0'' orbit, with one point (the centroid). If only two areal coordinates are equal, then we get ``type-1'' orbits with three points which lie on the medians of the triangle. Finally, if all three coordinates are different we get ``type-2'' orbits with six points.
A rule that uses $n_0$ type-0 orbits, $n_1$ type-1 orbits and $n_2$ type-2 orbits is called a rule of type $[n_0,n_1,n_2]$. The number of points for such a rule is
\begin{equation}
\label{eq:np}
n_p=n_0 + 3 n_1 + 6 n_2
\end{equation}
Due to the full symmetry employed, when integrating any of the quantities $L_1^i L_2^j L_3^{d-i-j}$ the sum in equation~(\ref{eq:gencub}) will only contain terms of the form
\begin{equation}
L_1^i L_2^j L_3^{d-i-j} +
L_1^i L_3^j L_2^{d-i-j} +
L_2^i L_1^j L_3^{d-i-j} +
L_2^i L_3^j L_1^{d-i-j} +
L_3^i L_1^j L_2^{d-i-j} +
L_3^i L_2^j L_1^{d-i-j}
\end{equation}
These terms are symmetric polynomials, and can therefore be written in terms of the elementary polynomials $\tilde{L}_1 = -(L_1+L_2+L_3)$, $\tilde{L}_2 = L_1 L_2 + L_2 L_3 + L_3 L_1$ and $\tilde{L}_3 = -L_1 L_2 L_3$. It is easily seen that only terms of the form $\tilde{L}_1^k \tilde{L}_2^l \tilde{L}_3^m$ with $k+2l+3m=d$ will be used. Indeed, since $\tilde{L}_1=-1$, only terms of the form $\tilde{L}_2^l \tilde{L}_3^m$ with $2l+3m \leq d$ are actually needed.
The cubature rule of order $d$ can therefore be obtained by requiring that equation~(\ref{eq:gencub}) is exact when the function $f$ is any of the terms $\tilde{L}_2^l \tilde{L}_3^m$ with $2 l + 3 m \leq d$. The number of non-negative solutions of $2 l + 3 m \leq d$ for $l$ and $m$, and therefore the number of equations that must be solved, is given by~\citep{oeisA001399}
\begin{equation}
\label{eq:numeqs}
n_e = 1+ \left\lfloor \frac{d^2+6 d}{12} \right\rfloor
\end{equation}
with $\lfloor x \rfloor$ denoting the largest integer that is less or equal to $x$. This is a significant reduction in the number of equations, approximately by a factor of 6 for large values of $d$, compared to the value $\bar{n}_e$ given in equation~(\ref{eq:ne_bar}) for the general case.
While areal coordinates allow for simple formulations of expressions on a generic triangle, they have the disadvantage of introducing three coordinates, instead of the two independent coordinates needed. For this reason, moment equations have generally been obtained using Cartesian or polar coordinates and referring to a specific triangle (exploiting the fact that all triangles are affine). In the fully symmetric case, however, we see that using areal coordinates we easily end up with only two ``coordinates'', the symmetric polynomials $\tilde{L}_2$ and $\tilde{L}_3$.
As will be seen shortly, the moment equations can be further simplified by using, instead of $\tilde{L}_2$ and $\tilde{L}_3$, the quantities
\begin{equation}
\label{eq:pqdef}
p = 1-3 \tilde{L}_2 \qquad \text{and}
\qquad
q = 1 - \frac{27}{2} \tilde{L}_3 - \frac{9}{2} \tilde{L}_2
\end{equation}
The cubature rule of order $d$ can therefore be obtained by requiring that equation~(\ref{eq:gencub}) is exact when the function $f$ is any of the terms $p^i q^j$ with $2 i + 3 j \leq d$ and $i,j \geq 0$.
The moment equations for a fully symmetric rule of degree~$d$ and type $[n_0,n_1,n_2]$ can thus be written as
\begin{equation}
\label{eq:eqs6gen}
\sum_{k=1}^{n_0} \bar{w}_{0,k} p_{0,k}^i q_{0,k}^j +
\sum_{k=1}^{n_1} 3 \bar{w}_{1,k} p_{1,k}^i q_{1,k}^j +
\sum_{k=1}^{n_2} 6 \bar{w}_{2,k} p_{2,k}^i q_{2,k}^j
= I_{i,j} \quad\text{with}\quad 2i+3j \leq d
\end{equation}
The right hand sides are the integrals
\begin{equation}
I_{i,j} = \frac{1}{A} \int_A p^i q^j \,\mathrm{d} A
\end{equation}
which can be easily computed analytically to give
\begin{equation}
I_{0,0} = 1,\: I_{1,0} = 1/4,\: I_{0,1} = 1/10,\: I_{2,0} = 1/10,\: I_{1,1} = 2/35,\: I_{3,0} = 29/560,\: I_{0,2} = 7/160,\: I_{2,1} = 1/28, \ldots
\end{equation}
The main advantage of using the quantities $p$ and $q$ is that for type-1 orbits we can introduce a new variable $u$ so that $p=u^2$ and $q=u^3$ and therefore $p^i q^j = u^{2 i + 3 j}$, while for the type-0 orbit $p=q=0$. Setting $w_0 = \bar{w}_{0,1}$, $v_k = 3 \bar{w}_{1,k}$ and $w_k = 6 \bar{w}_{2,k}$, after some computations, the moment equations are finally written as
\begin{subequations}
\label{eq:momeqs}
\begin{gather}
\label{eq:momeq1}
w_0 + \sum_{k=1}^{n_1} v_k + \sum_{k=1}^{n_2} w_k = I_{0,0}\\
\label{eq:momeq3}
\sum_{k=1}^{n_1} v_k u_k^{2i+3j} + \sum_{k=1}^{n_2} w_k p_k^i q_k^j = I_{i,j}
\quad\text{with}\quad 2i+3j \leq d, j \leq 1\\
\label{eq:momeq6}
\sum_{k=1}^{n_2} w_k (p_k^3-q_k^2) p_k^i q_k^j = I_{i+3,j} - I_{i,j+2}
\quad\text{with}\quad 2i+3j \leq d-6
\end{gather}
\end{subequations}
where in equation~(\ref{eq:momeq1}) we set $w_0=0$ if $n_0=0$.
For both $d=0$ and $d=1$ the only moment equation is~(\ref{eq:momeq1}). This means that any (fully symmetric) rule exact for $d=0$ will also be exact for $d=1$, thus there are no rules of degree 0. For this reason in the following we always assume that $d \geq 1$.
\subsection{Consistency conditions}
\label{sec:conscond}
To set up the moment of equations for a rule of degree $d$, it is first necessary to determine the type of the rule, i.e. the number of orbits of each type.
The moment equations~(\ref{eq:momeqs}) form a system of $n_e$ equations in $n_v$ variables, where $n_e$ is given in equation~(\ref{eq:numeqs}) while $n_v = n_0 + 2 n_1 + 3 n_2$. Similarly, the subsystem~(\ref{eq:momeq6}) has $n_e-d$ equations and $3 n_2$ variables.
We assume that both the system~(\ref{eq:momeqs}) and its subsystem~(\ref{eq:momeq6}) are inconsistent (i.e.\ have no solutions) if and only if they are overdetermined (i.e. have more equations than variables). This assumption, together with the fact that there may be at most one type~0 orbit, yields the following \emph{consistency conditions}
\begin{subequations}
\begin{gather}
3 n_2 \geq n_e - d \\
3 n_2 + 2 n_1 + n_0 \geq n_e \\
n_0 \leq 1
\end{gather}
\end{subequations}
which must be satisfied to obtain a solution of the moment equations, and thus they restrict the choice of the rule type. For a given degree $d$, a minimal-point rule is sought, that is a rule that satisfies the consistency conditions with the lowest total number of points, as given by equation~(\ref{eq:np}). This yields
\begin{equation}
n_2 = \big\lfloor (n_e - d + 2)/3 \big\rfloor
,\qquad
n_1 = \big\lfloor (n_e - 3 n_2)/2 \big\rfloor
,\qquad
n_0 = n_e - 3 n_2 - 2 n_1
\end{equation}
It is conceivable that a rule that violates the consistency conditions may lead to a system of moment equations that, although overdetermined, has solutions. These so-called \emph{fortuitous} rules have great theoretical interest, as well as practical interest in the case where they have fewer integration points compared to the minimal-point rules described above. No fortuitous rules are encountered in the present paper, however, nor in the available literature on cubature rules on the triangle.
The system of moment equations~(\ref{eq:momeqs}) can be inconsistent, zero-dimensional or positive-dimensional (with zero solutions, a finite number of solutions or infinite solutions respectively). We use here the same terms to identify the corresponding rule types and individual rules, thus we have inconsistent rule types, which yield no rules, zero-dimensional rule types, which yield a finite number of zero-dimensional rules, and positive-dimensional rule types which yield an infinite number of positive dimensional rules. In the case of positive-dimensional rule types, the analytical solution can be expressed using a number of free parameters.
\subsection{Advantages of the suggested form of the moment equations}
The development of the method given in Sections~\ref{sec:momeqs} and~\ref{sec:conscond} to formulate the moment equations using symmetric polynomials follows in some main points the classic one presented by \citet{LynessJespersen1975}. It has, however, the obvious benefit of providing polynomial moment equations, while~\citep{LynessJespersen1975} also uses cosines. In this, the present method is similar to the one presented by \citet{WandzuraXiao2003}.
All three methods are equivalent, in that they yield the same rules. Indeed, it is relatively easy to pass from one method to the other: setting $p_i = r_i^2$, $q_i = r_i^3 \cos 3 \alpha_i$ and $u_i = r_i$ in equations~(\ref{eq:momeqs}) yields the moment equations in~\citep{LynessJespersen1975}, while it is easily seen that, for the triangle used in~\citep{WandzuraXiao2003}, $p$ and $q$ are equal to the invariants $x^2+y^2$ and $x^3-3 x y^2$.
The present method is arguably simpler and more intuitive in its formulation, while it provides simpler formulas. Additionally, this method is elegantly formulated without reference to a specific triangle. From a practical point of view, however, the main advantage is that the resulting polynomial equations are of significantly lower degree than those provided by the other methods, for example the maximum degree of equations~(\ref{eq:momeq6}) is $\lfloor d/2 \rfloor + 1$ instead of $d+1$. This is especially important when solving the equations analytically.
\section{Analytical solution of the moment equations}
\label{sec:analytsol}
\subsection{The usefulness of analytical solutions}
Except for some trivial low-degree rules, the moment equations are generally solved numerically, e.g.\ using a multivariate Newton-Raphson solver. The cubature rule is then given as a table of integration point coordinates and weights, expressed as floating point approximations of a given precision. This numerical approximation of the cubature rule is the one actually required when using the rule in applications.
Numerical methods have the advantage of being able to provide cubature rules of high degree~\citep[see e.g.][]{XiaoGimbutas2010}. Convergence of the method to a solution is not guaranteed, however, as it most often depends on the selection of an appropriate ``initial guess'' required by the solver. This means that inability to obtain a solution does not prove that the solution does not exist. Additionally, when a solution is obtained numerically, no information is obtained regarding the existence of other solutions.
For this reason, in this paper we investigate the analytical solution of the moment equations, in order to obtain a definitive answer regarding the different cubature rules for a given degree and type.
There exist algorithms for solving analytically arbitrary systems of polynomial equations, for example using Gr\"obner bases (see~\citep{Lazard2009} for an informal overview of the state of the art). Unfortunately, their requirements in both computer memory and computation time are such that in practice they fail to provide a solution even for rules of relatively low degree. To obtain solutions for higher degrees, it is therefore necessary to exploit as much as possible the structure of the moment equations.
An interesting alternative to the analytical solution of the moment equations is to use homotopy continuation methods to compute \emph{all} solutions of the system numerically~\citep{Verschelde1999}. This is however clearly beyond the scope of the present paper.
\subsection{Solution strategy}
The subsystems~(\ref{eq:momeq1}), (\ref{eq:momeq3}) and~(\ref{eq:momeq6}) have respectively $1$, $d-1$ and $n_e-d$ equations. The weight $w_0$ (if it is non-zero) appears only in equation~(\ref{eq:momeq1}) while the variables $v_k$ and $u_k$ appear only in equations~(\ref{eq:momeq1}) and (\ref{eq:momeq3}).
Consider first the case of a rule with a type-0 orbit ($n_0=1$). Equation~(\ref{eq:momeq1}) is then just used to determine $w_0$ when all other weights have been calculated. The weights $v_k$ of type-1 orbits can be eliminated from equations~(\ref{eq:momeq3}), as described in~\citep[pp.~771--773]{RabinowitzRichter1969} for cubature rules on other regions, to obtain the (linear in the symmetric polynomials $\tilde{u}_k$) system of equations
\begin{equation}
\label{eq:momeq3sym}
\sum_{k=0}^{n_1} J_{i-k} \tilde{u}_k = 0, \qquad i=n_1+2, \ldots, d
\end{equation}
where
\begin{equation}
\label{eq:Jdef}
J_i = \begin{cases}
\displaystyle{ I_{j,0} - \sum_{k=1}^{n_2} w_k p_k^j } & \text{if $i=2j$} \\[2em]
\displaystyle{ I_{j,1} - \sum_{k=1}^{n_2} w_k p_k^j q_k } & \text{if $i=2j+3$}
\end{cases}
\end{equation}
The system~(\ref{eq:momeq3sym}) has $n_1$ unknowns $\tilde{u}_k$ (since $\tilde{u}_0 = 1$) and $d-n_1-1$ equations. If $n_1=(d-1)/2$ then equations~(\ref{eq:momeq6}) are sufficient to evaluate the variables $w_k$, $p_k$ and $q_k$ of type-2 orbits, and then equations~(\ref{eq:momeq3sym}), (\ref{eq:momeq3}) and~(\ref{eq:momeq1}) yield in turn the values of $\tilde{u}_k$, $v_k$ and $w_0$. The same happens if $n_1>(d-1)/2$, but in this case the system is positive-dimensional and some of the $\tilde{u}_k$ remain as free parameters in the solution. Finally, if $n_1<(d-1)/2$ then obtaining a solution is more difficult, since to evaluate $w_k$, $p_k$ and $q_k$ we need not only equations~(\ref{eq:momeq6}) but also the equations that remain after eliminating $\tilde{u}_k$ from~(\ref{eq:momeq3sym}).
When the type-0 orbit is not used ($n_0=0$), it is generally easier to introduce an additional equation
\begin{equation}
\label{eq:momeqextra}
\sum_{k=1}^{n_1} v_k u_k = J_1
\end{equation}
where $J_1$ is an unknown quantity, which is not defined by~(\ref{eq:Jdef}). Eliminating the weights $v_k$ from equations~(\ref{eq:momeq1}), (\ref{eq:momeq3}) and~(\ref{eq:momeqextra}) leads to a system of equations like~(\ref{eq:momeq3sym}), only that the index $i$ is now in the range $i=n_1, \ldots, d$ and $J_1$ is an additional unknown that must be eliminated.
In all cases, equations~(\ref{eq:momeq6}) must be solved, possibly together with the equations that remain after eliminating $\tilde{u}_k$ from~(\ref{eq:momeq3sym}). Unfortunately, no easy way has been found to simplify these equations as we did to derive the system~(\ref{eq:momeq3sym}). The use of symmetric polynomials can, however, again lead to somehow simpler expressions.
\subsection{Permutation invariance of the orbits}
\label{sec:orbinv}
In Section~\ref{sec:formulating} we exploited the fact that the cubature rule is invariant with respect to a permutation of the integration points within a given orbit, and expressed this invariance using symmetric polynomials.
Another obvious property of the cubature rules, which however has received much less attention in the literature, is their invariance with respect to permutation of orbits of the same type. This is reflected in the fact that the moment equations~(\ref{eq:momeqs}) are polynomials which are ``symmetric'' (i.e.\ invariant with respect to permutation) in the pairs ($u_k$,$v_k$) and in the triplets ($p_k$, $q_k$, $w_k$). This can be seen from the system~(\ref{eq:momeq3sym}) where, having eliminated the $v_k$, the resulting polynomials are symmetric in the $u_k$ and have thus been expressed in terms of the elementary symmetric polynomials $\tilde{u}_k$. In a similar way, eliminating $q_k$ and $w_k$ allows us to express the moment equations in terms of the symmetric polynomials $\tilde{p}_k$.
The system that results by eliminating the $v_k$, $q_k$ and $w_k$ from the moment equations and expressing the results in terms of the $\tilde{u}_k$ and $\tilde{p}_k$ is in most cases more complicated than the moment equations. It has however fewer variables, and it leads to a much simpler expression for the solution, when such a solution is actually found.
Indeed, one important advantage of expressing the moment equations in terms of symmetric polynomials is that the number of solutions of the system is equal to the number of different cubature rules that can be obtained. Consider for example the degree-4 $[0,2,0]$ rule, for which~\citet{LynessJespersen1975} mention that, in the present notation, $u_1$ and $u_2$ are the roots of $15 x^4 + 20 x^3 - 30 x^2 + 4$. This does not mean, however, than any combination of the roots is a valid solution for $u_1$ and $u_2$, indeed only two pairs of solutions give a cubature rule. In terms of symmetric polynomials, on the other hand, the solution is obtained by solving the equations $3\tilde{u}_1^2-4\tilde{u}_1-2 = 0$ and $5\tilde{u}_2+2\tilde{u}_1+2 = 0$, where it is seen that two different rules are obtained, one for each solution of the system.
It is worth considering that even when solving the moment equations numerically, considering the invariance with respect to permutation of orbits of the same type can have a significant effect on the solution method. As an example, there is only one degree-15 [1,7,4] rule. The system~(\ref{eq:momeq6}) however has $4!=24$ solutions, while if we were to solve all equations~(\ref{eq:momeqs}) together we would have $7!4!=120960$ solutions. It is thus conceivable that an iterative numerical solution algorithm may fail to converge by being ``attracted'' in turn by different solutions.
\subsection{Solution quality}
Once a cubature rule is determined by solving the moment equations, the sign of the weights and the position of the integration points is examined, to determine the \emph{quality} of the solution. The quality is described using a two-letter label: the first letter is P if all weights are positive and N if at least one weight is negative, while the second letter is I if all points are inside the triangle, O if there is at least one point outside the triangle, and B if no points are outside the triangle but at least one is on the boundary of the triangle. The following qualities are therefore encountered: PI, NI, PB, NB, PO, NO.
In all the above cases, the coordinates and weights of the integration points are considered to be real. Though it is well-known that complex solutions may exist, these are not taken into account, since a cubature rule with complex-valued coordinates of the integration points would be of little, if any, use. Moreover, the moment equations are usually solved using numerical methods that only return real solutions, as these methods perform significantly better than methods that could return complex solutions.
On the other hand, when obtaining the solutions analytically it costs nothing to also consider complex solutions. For this reason, we expand the above definition of the quality of cubature rules by setting the first letter of the label to C if at least one weight is complex-valued and by setting the second letter of the label to C if at least one integration point has complex coordinates. Interestingly, while it is not possible to have complex weights without complex coordinates, it is possible to have real weights with complex coordinates. The following three additional qualities are therefore obtained: CC, PC, NC.
Including complex solutions allows us to make the distinction between moment equations that have no solution and those that have solutions, even though they may all be complex. Considering as an example a degree-15 rule, there are no solutions for type $[0,7,4]$ (which does not satisfy the consistency conditions), while there is a single complex (NC) solution for type $[1,7,4]$ (which satisfies the consistency conditions). It is generally expected that all types satisfying the consistency conditions will yield at least one solution, but with complex solutions appearing with increasing frequency as the degree of the rule increases.
Although we compute all solutions, independently of their quality, in most applications we need rules of PI (or at most NI) quality. For this reason, if a minimal-point rule does not yield any PI rules, we investigate rules with increasingly more points until a rule is found that has a PI solution. When considering rules with additional points, it is possible to have rules with the same degree and number of points, but different type and different number of free parameters appearing in the solution.
Consider for example the degree-7 rules. The minimal-point rule $[1,2,1]$ has 13 points and the best quality achievable with it is NI. Increasing the number of points, we get either a $[0,3,1]$ or a $[0,1,2]$ rule, both with 15 points, where the first has one free parameter while the second has none. In this case, where both types can yield PI rules, we prefer the zero-dimensional one as it has more type-2 orbits, so less integration points are restricted to be located on the medians.
In general, among rules with the same number of points and the same quality, we would prefer those with more type-2 orbits and thus less free parameters. The presence of free parameters in the solution of the moment equations, on the other hand, allows for much greater flexibility in obtaining a rule of PI quality. Moreover, the use of more type-1 orbits leads to simpler moment equations, which are easier to solve analytically.
Note that the numerical, iterative solution of the moment equations for positive-dimensional rules \citep[see e.g.][]{WandzuraXiao2003} yields only one of the infinite solutions. Though it is possible to consider numerically the variation of the solution with the variation of a parameter \citep[see e.g.][]{BerntsenEspelid1990}, analytical solutions are much more powerful in studying parametrically positive-dimensional cubature rules and their quality. The study and presentation of such rules, however, requires a much more extensive discussion which goes well beyond the scope of the present paper. For this reason, in Section~\ref{sec:results} we only present results for zero-dimensional cubature rules.
\section{Results and discussion}
\label{sec:results}
Using the method described in Sections~\ref{sec:formulating} and~\ref{sec:analytsol} we compute here analytically cubature rules for degree up to~15. As described in Section~\ref{sec:orbinv}, the permutation invariance of the orbits should be exploited to express the moment equations~(\ref{eq:momeqs}) in a form more suitable for analytical solution, for example in terms of the symmetric polynomials $\tilde{u}_k$ and $\tilde{p}_k$. This has been achieved for each degree and rule type in a heuristic way, which involved (for higher degrees) extensive calculations until the initial system was transformed into a new one, solvable (on the available hardware and software) using Gr\"obner bases. The actual calculations performed in each case are obviously too lengthy to be written out here. Indeed, in the non-trivial cases, the analytical solution itself becomes too long, as is already apparent in Appendix~\ref{sec:app_analyt} for the degree-6 rule.
Table~\ref{tab:rulesummary} gives a summary of the properties of all cubature rules thus computed. As already mentioned, we only consider zero-dimensional rules. We calculate for each degree the minimal-point rules and, if none of these are of quality PI, we calculate additional rule types with more points until a rule with PI quality is found (except for $d=15$ where additional rules were not computed). Appendix~\ref{sec:app_analyt} provides analytical expressions for evaluating some of the cubature rules, while Appendix~\ref{sec:app_numer} provides numerical values for new rules of PI or NI quality.
\begin{table}
\centering
\caption{Summary of the properties of all computed rules}\label{tab:rulesummary}
\begin{tabular}{rcrrrrrrrrrr}
\toprule
degree & type & points & solutions & PI & NI & PB & PO & NO & PC & NC & CC \\
\midrule
1 & $[1,0,0]$ & 1 & 1 & 1 & -- & -- & -- & -- & -- & -- & -- \\
2 & $[0,1,0]$ & 3 & 2 & 1 & -- & 1 & -- & -- & -- & -- & -- \\
3 & $[1,1,0]$ & 4 & 1 & -- & 1 & -- & -- & -- & -- & -- & -- \\
& $[0,0,1]$ & 6 & 1 & 1 & -- & -- & -- & -- & -- & -- & -- \\
4 & $[0,2,0]$ & 6 & 2 & 1 & -- & -- & 1 & -- & -- & -- & -- \\
5 & $[1,2,0]$ & 7 & 1 & 1 & -- & -- & -- & -- & -- & -- & -- \\
6 & $[0,2,1]$ & 12 & 6 & 2 & -- & -- & 2 & -- & -- & -- & 2 \\
7 & $[1,2,1]$ & 13 & 4 & -- & 1 & -- & 1 & -- & -- & -- & 2 \\
& $[0,1,2]$ & 15 & 4 & 2 & -- & -- & -- & -- & -- & -- & 2 \\
8 & $[1,3,1]$ & 16 & 2 & 1 & 1 & -- & -- & -- & -- & -- & -- \\
9 & $[1,4,1]$ & 19 & 1 & 1 & -- & -- & -- & -- & -- & -- & -- \\
10 & $[0,4,2]$ & 24 & 14 & -- & -- & -- & 4 & 1 & -- & -- & 9 \\
& $[1,2,3]$ & 25 & 15 & 4 & -- & -- & -- & 2 & -- & 3 & 6 \\
11 & $[0,5,2]$ & 27 & 6 & -- & -- & -- & 1 & -- & -- & 2 & 3 \\
& $[1,3,3]$ & 28 & 23 & -- & 2 & -- & 5 & 3 & 2 & 4 & 7 \\
& $[0,2,4]$ & 30 & 34 & 4 & -- & -- & 1 & 1 & 4 & 2 & 22 \\
12 & $[0,5,3]$ & 33 & 24 & 2 & 1 & -- & -- & -- & -- & -- & 21 \\
13 & $[0,6,3]$ & 36 & 8 & -- & -- & -- & -- & 1 & -- & 1 & 6 \\
& $[1,4,4]$ & 37 & 54 & 2 & 3 & -- & 4 & 5 & 2 & 8 & 30 \\
14 & $[0,6,4]$ & 42 & 38 & 1 & -- & -- & 3 & 3 & -- & 3 & 28 \\
15 & $[1,7,4]$ & 46 & 1 & -- & -- & -- & -- & -- & -- & 1 & -- \\
\bottomrule
\end{tabular}
\end{table}
The only case where three rule types must be computed to obtain PI quality is $d=11$. This is therefore the only case (for $d<15$) where a positive-dimensional rule of PI quality (type $[1,5,2]$ with 28 points) has less points than the best possible zero-dimensional rule of the same quality (type $[0,2,4]$ with 30 points).
\footnote{Type $[1,5,2]$ fully symmetric PI rules exist, and are easy to obtain using the method presented in this paper, yet no such rule was encountered in the literature. \citet{LynessJespersen1975} present a $[1,5,2]$ PB rule, \citet{WandzuraXiao2003} compute what is most probably a $[1,5,2]$ PI rule but do not present it, while the 28-point PB rule given by \Citet{Taylor2007} and the 27-point PI rule given by \citet{Taylor2008} are asymmetric rules.}
Where NI rules are acceptable, the two $[1,3,3]$ rules can be used, since they have the same number of points as the $[1,5,2]$ rules and have less points on the medians. The second $[1,3,3]$ NI rule given in Appendix~\ref{sec:app_numer} is then to be preferred as it has a small negative weight for a single point, while the first one has a large negative weight for three points.
The results summarised in Table~\ref{tab:rulesummary} confirm the general expectation that as the rule degree increases the number of solutions will increase, though with most solutions being complex ones. This is not always the case, however, as evidenced by the existence of a single $[1,7,4]$ rule for $d=15$. It is thus clear that it is not possible to detect in these results a specific pattern in the number of solutions, the number of real solutions or the number of PI (or NI) solutions.
An interesting side effect of computing analytically the cubature rules is that we can prove the non-existence of specific fortuitous rules. Consider for example the case of degree-10 rules where we compute the 24-point $[0,4,2]$ rule. If there were a fortuitous rule with two type-2 orbits and less than 24 points, then the $[0,4,2]$ rule should be positive-dimensional in order to depend on some parameters which, for specific values, would yield the fortuitous rule. Computing analytically the $[0,4,2]$ rule, however, shows that it is zero-dimensional, as expected. Similarly, since the $[1,2,3]$ rule is zero-dimensional, there exist no fortuitous rules with three type-2 orbits and less than 25 points. Since it is easily shown that for degree 10 no rules exist with one or zero type-2 orbits and also that no $[0,0,4]$ rules exist (which would have 24 points) we see that there are no fortuitous degree-10 rules with 24 points or less. Similar tests can be performed for all other rule degrees considered here.
A list of numerical values for all computed rules, independently of their quality, is provided together with this paper as supplemental material. This list includes the zero-dimensional rules found in~\citep{Stroud1971, Cowper1973, LynessJespersen1975, LaursenGellert1978, Dunavant1985}. An interesting property of some rules of bad quality (i.e. neither PI nor NI) is that the orbits that have points outside the triangle or with complex coordinates have a much smaller weight (in absolute value). This is the case for example for $d=11$ and the fourth $[1,3,3]$ NC rule, or for $d=15$ and the $[1,7,4]$ NC rule. These rules, together with a node elimination algorithm \citep{XiaoGimbutas2010}, could possibly be used to derive cubature rules that are not fully symmetric with fewer points than the fully symmetric ones.
\section{Conclusions}
In this paper we have used symmetric polynomials to express the double invariance inherent in fully symmetric cubature rules in the triangle (invariance with respect to permutation of points within an orbit and with respect to permutation of orbits of the same type). This has allowed us to formulate the moment equations in such a way that analytical solutions have been derived for zero-dimensional rules of degree up to 15.
A few new rules of good quality have been thus derived and are given in Appendix~\ref{sec:app_numer}. Additionally, the analytical solutions ensure that all possible rules of a given type and degree were computed, independently of their quality. This allows us, for example, to prove that indeed no rules of good quality (PI or even NI) exist for some cases where no such rules were encountered in the literature.
Though only zero-dimensional rules have been computed here, the proposed analytical approach is also well-suited for the thorough study of positive-dimensional rules. In this case, however, an additional difficulty lies in finding intuitive and useful ways to present the (infinite) solutions and their properties.
In all cases, combining a better understanding of the structure of the moment equations together with better-performing algorithms, software implementations and hardware platforms, should allow determining rules of increasingly high degree.
\section*{Acknowledgements}
The research leading to these results has received funding from the
European Research Council under the European Community's Seventh
Framework Programme (FP7/2007--2013) / ERC grant agreement n\textsuperscript{o} 228051 [MEDIGRA].
|
\section{Introduction}\label{sec-int}
The study of probability-measure-preserving actions of discrete countable groups through the associated orbit equivalence relations has been making progress since the seminal works \cite{cfw}, \cite{dye1}, \cite{dye2} and \cite{ow} on amenable groups.
We recommend the reader to consult \cite{furman-survey}, \cite{gab}, \cite{popa-survey}, \cite{shalom-survey} and \cite{vaes-survey} for recent development and related topics.
In this paper, we focus on actions of non-amenable Baumslag-Solitar groups, and introduce their invariants under orbit equivalence.
For two non-zero integers $p$, $q$, the {\it Baumslag-Solitar group} ${\rm BS}(p, q)$ is defined as the group with the presentation
\[{\rm BS}(p, q)=\langle \, a, t\mid ta^pt^{-1}=a^q\,\rangle.\]
After the appearance in \cite{bs} as a simple example of a non-Hopfian finitely presented group for certain $p$ and $q$, this group attracts attention in combinatorial group theory.
The isomorphism problem for these groups is solved by Moldavanski\u{\i} \cite{mold}, who shows that ${\rm BS}(p, q)$ and ${\rm BS}(r, s)$ are isomorphic if and only if there exists an integer $\varepsilon \in \{ \pm 1\}$ such that either $(p, q)=(\varepsilon r, \varepsilon s)$ or $(p, q)=(\varepsilon s, \varepsilon r)$.
The group ${\rm BS}(p, q)$ is amenable if and only if either $|p|=1$ or $|q|=1$.
We therefore assume $2\leq |p|\leq |q|$ throughout the paper.
A discrete group is assumed countable unless otherwise mentioned.
We mean by an {\it f.f.m.p}.\ action of a discrete group $G$ a measure-preserving Borel action of $G$ on a standard finite measure space $(X, \mu)$ such that the stabilizer of a.e.\ $x\in X$ in $G$ is trivial, where ``f.f.m.p." stands for ``essentially free and finite-measure-preserving".
The following equivalence relations among such actions are of our interest.
\begin{defn}
Let $G\c (X, \mu)$ and $H\c (Y, \nu)$ be ergodic f.f.m.p.\ actions of discrete groups.
We say that these actions are {\it orbit equivalent (OE)} if there exists a Borel isomorphism $f$ from a conull Borel subset of $X$ onto a conull Borel subset of $Y$ such that $f_*\mu$ and $\nu$ are equivalent and we have $f(Gx)=Hf(x)$ for a.e.\ $x\in X$.
More generally, we say that the actions $G\c (X, \mu)$ and $H\c (Y, \nu)$ are {\it weakly orbit equivalent (WOE)} if there exists a Borel isomorphism $f$ from a Borel subset $A$ of $X$ with $\mu(A)>0$ onto a Borel subset $B$ of $Y$ with $\nu(B)>0$ such that $f_*(\mu|_A)$ and $\nu|_B$ are equivalent and we have $f(Gx\cap A)=Hf(x)\cap B$ for a.e.\ $x\in A$.
\end{defn}
We put $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$ and define a homomorphism $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$, called the {\it modular homomorphism} of $\Gamma$, by setting $\mathfrak{m}(a)=1$ and $\mathfrak{m}(t)=|q/p|$.
Let $\Gamma \c (X, \mu)$ be an ergodic f.f.m.p.\ action.
We now introduce the flow associated with this action, which is shown to be an invariant under orbit equivalence.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ be the ergodic decomposition for the action of $\ker \mathfrak{m}$ on $(X, \mu)$.
We have the canonical ergodic measure-preserving action of $\mathfrak{m}(\Gamma)$ on $(Z, \xi)$.
The {\it flow associated with} the action $\Gamma \c (X, \mu)$ is defined as the action of $\mathbb{R}$ induced from the action of $\log \circ \mathfrak{m}(\Gamma)$ on $(Z, \xi)$ through the isomorphism $\log \colon \mathbb{R}_+^{\times} \to \mathbb{R}$.
We refer to Remark \ref{rem-type} for a detailed description of this flow.
\begin{thm}\label{thm-flow}
We set $\Gamma ={\rm BS}(p, q)$ and $\Lambda ={\rm BS}(r, s)$ with $2\leq |p|\leq |q|$ and $2\leq |r|\leq |s|$.
Let $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ be ergodic f.f.m.p.\ actions.
If these two actions are WOE, then the flows associated with them are isomorphic.
\end{thm}
For a locally compact group $G$, two measure-preserving actions $G\c (Z, \xi)$ and $G\c (W, \omega)$ on measure spaces are called {\it isomorphic} if there exists a measurable isomorphism $f$ from a conull measurable subset of $Z$ onto a conull measurable subset of $W$ such that $f_*\xi$ and $\omega$ are equivalent and for any $g\in G$, we have $f(gz)=gf(z)$ for a.e.\ $z\in Z$.
In Corollary \ref{cor-conj}, we obtain a refinement of Theorem \ref{thm-flow} under an additional assumption.
Theorem \ref{thm-flow} is strongly inspired by the theory of the flow associated with an ergodic transformation of type ${\rm III}$, summarized in \cite{ho}.
This flow is defined as the Mackey range of the Radon-Nikodym cocycle for the transformation.
It is notable that the isomorphism class of the associated flow is a complete invariant of ergodic single transformations of type ${\rm III}$ under orbit equivalence.
This is due to Hamachi-Oka-Osikawa \cite{hoo} and Krieger \cite{krieger}, \cite{krieger-flow}.
Theorem \ref{thm-flow} is indeed a consequence of a rigidity result on the composition of the cocycle associated with the WOE and the homomorphism $\mathfrak{m}$, which is formulated in terms of measure equivalence in Theorem \ref{thm-r}.
In Section \ref{sec-var}, we prove a variant of Furman's theorem on construction of a representation of a group which is measure equivalent to a given group.
Combining these results, we obtain the following:
\begin{thm}\label{thm-z}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|<|q|$.
Suppose that an ergodic f.f.m.p.\ action $\Gamma \c (X, \mu)$ is WOE to a weakly mixing f.f.m.p.\ action $\Delta \c (Y, \nu)$ of a discrete group $\Delta$.
Then there exists a homomorphism from $\Delta$ onto $\mathbb{Z}$.
\end{thm}
Putting $\Gamma ={\rm BS}(p, q)$, we mean by an {\it elliptic} subgroup of $\Gamma$ a subgroup contained in the group generated by $a$ or in its conjugate in $\Gamma$.
Under the assumption that any non-trivial elliptic subgroup of $\Gamma$ acts ergodically, the parameters $p$, $q$ are shown to be invariant under WOE, while the associated flow is isomorphic to the action of $\mathbb{R}$ on $\mathbb{R} /(\log |q/p|)\mathbb{Z}$ by addition, and remembers at most the modulus $|q/p|$.
\begin{thm}\label{thm-woe}
We set $\Gamma ={\rm BS}(p, q)$ and $\Lambda ={\rm BS}(r, s)$ with $2\leq |p|<|q|$ and $2\leq |r|<|s|$.
Suppose that $q$ is not a multiple of $p$ and that $s$ is not a multiple of $r$.
Let $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ be f.f.m.p.\ actions such that any non-trivial elliptic subgroup of $\Gamma$ acts ergodically and so does any non-trivial elliptic subgroup of $\Lambda$.
If the two actions $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ are WOE, then there exists an integer $\varepsilon \in \{ \pm 1\}$ with $(p, q)=(\varepsilon r, \varepsilon s)$, that is, $\Gamma$ and $\Lambda$ are isomorphic.
\end{thm}
The assumption on the actions of non-trivial elliptic subgroups of $\Gamma$ and $\Lambda$ can be relaxed.
We refer to Theorem \ref{thm-mer} for a more general assertion.
In Section \ref{sec-oe}, when $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$, we construct two f.f.m.p.\ actions of $\Gamma$ which are WOE, but not conjugate.
The restrictions of the two actions to any non-trivial elliptic subgroup of $\Gamma$ are shown to be ergodic.
It is therefore impossible to obtain conjugacy of the two actions under the assumption in Theorem \ref{thm-woe}.
As for Bernoulli shifts of Baumslag-Solitar groups, Popa's cocycle superrigidity theorem \cite{popa-gap} shows orbit equivalence rigidity of them (see Remark \ref{rem-ber}).
We now turn our attention to measure equivalence (ME).
This is an equivalence relation between discrete groups, introduced by Gromov \cite{gromov-as-inv} (see Definition \ref{defn-me}).
It is known that two discrete groups $G$ and $H$ are ME if and only if there exists an ergodic f.f.m.p.\ action of $G$ which is WOE to an ergodic f.f.m.p.\ action of $H$.
\begin{thm}\label{thm-bs-hyp}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|<|q|$.
Let $\Delta$ be a discrete group having an infinite amenable normal subgroup $N$.
Suppose that the quotient $\Delta /N$ is a non-elementarily word-hyperbolic group.
Then $\Gamma$ and $\Delta$ are not ME.
\end{thm}
For any integers $r$, $s$ with $2\leq |r|=|s|$, the group ${\rm BS}(r, s)$ contains a finite index subgroup isomorphic to the direct product of $\mathbb{Z}$ with a non-abelian free group of finite rank.
It follows from Theorem \ref{thm-bs-hyp} that the group ${\rm BS}(p, q)$ with $2\leq |p|<|q|$ is not ME to ${\rm BS}(r, s)$.
The basic question asking whether the groups ${\rm BS}(p, q)$ with different parameters $p$, $q$ are ME or not remains unsolved.
This paper is organized as follows.
In Section \ref{sec-bs}, we review basic properties of Baumslag-Solitar groups and the Bass-Serre trees associated with them.
In Section \ref{sec-dmg}, starting with terminology of discrete measured groupoids, we introduce several notions related to them, index, local index, normality, quasi-normality and quotient.
In Section \ref{sec-var}, we review the aforementioned Furman's theorem and prove its variant.
In Section \ref{sec-ell}, given a finite-measure-preserving action of a Baumslag-Solitar group, we present an algebraic and sufficient condition for a subgroupoid of the associated groupoid to be elliptic.
Elliptic subgroupoids are shown to be preserved under an isomorphism between the groupoids associated with actions of Baumslag-Solitar groups.
In Section \ref{sec-mod}, we introduce the modular cocycle of Radon-Nikodym type and the local-index cocycle.
They are defined for the pair of a discrete measured groupoid and its quasi-normal subgroupoid.
If the pair comes from an action of a Baumslag-Solitar group and its restriction to an elliptic subgroup, then those two cocycles are related to the modular homomorphism.
In Section \ref{sec-mackey}, Theorems \ref{thm-flow}, \ref{thm-z} and \ref{thm-bs-hyp} are proved by using those cocycles.
In Section \ref{sec-red}, Theorem \ref{thm-woe} is proved.
In Section \ref{sec-oe}, we discuss f.f.m.p.\ actions of a Baumslag-Solitar group which are WOE, but not conjugate.
In Appendix \ref{app}, we observe basic properties of an ergodic probability-measure-preserving action of a Baumslag-Solitar group, focusing on ergodicity of the action of an elliptic subgroup.
Those properties are used to relax ergodicity assumptions in theorems.
In Appendix \ref{sec-exotic}, we show that the groupoid associated with a certain action of a Baumslag-Solitar group has an amenable normal subgroupoid of infinite type.
We also show that the quotient groupoid is of type ${\rm III}$, and clarify relationship with the associated flow.
\section{Baumslag-Solitar groups}\label{sec-bs}
Contents of this section are discussed from a more general viewpoint in \cite[Section 2]{levitt}.
For a group $G$ and an element $g$ of $G$, let $\langle g \rangle$ denote the cyclic subgroup of $G$ generated by $g$.
Let $p$ and $q$ be integers with $2\leq |p|\leq |q|$, and set
\begin{equation*}\tag{$\star$}\label{pre}
\Gamma ={\rm BS}(p, q)=\langle\, a, t\mid ta^pt^{-1}=a^q\,\rangle.
\end{equation*}
The group $\Gamma$ is the HNN extension of the infinite cyclic group $\langle a\rangle$ relative to the isomorphism from $\langle a^p\rangle$ onto $\langle a^q\rangle$ sending $a^p$ to $a^q$.
Let $T=T_{\Gamma}$ denote the Bass-Serre tree associated with this HNN extension.
The set of vertices of $T$, denoted by $V(T)$, is defined to be $\Gamma /\langle a\rangle$.
The set of edges of $T$, denoted by $E(T)$, is defined to be $\Gamma/ \langle a^q\rangle$, and for each $\gamma \in \Gamma$, the edge corresponding to the coset $\gamma \langle a^q\rangle$ joins the two vertices corresponding to the cosets $\gamma \langle a\rangle$ and $\gamma t\langle a\rangle$.
We orient this edge so that the origin is the vertex corresponding to $\gamma \langle a\rangle$.
The group $\Gamma$ then acts on $T$ by orientation-preserving simplicial automorphisms.
The action of $\Gamma$ on $V(T)$ and that on $E(T)$ are both transitive.
Let ${\rm Aut}(T)$ denote the group of orientation-preserving simplicial automorphisms of $T$ equipped with the standard Borel structure induced by the pointwise convergence topology.
Unless otherwise stated, we mean by the {\it Bass-Serre tree associated with} $\Gamma$ the oriented simplicial tree defined above for a fixed presentation of $\Gamma$ of the form (\ref{pre}).
We refer to \cite{serre} for the Bass-Serre theory.
Let $E_+$ denote the set of oriented edges of $T$ with the orientation defined above.
Let $E_-$ denote the set of oriented edges of $T$ consisting of the inverses of edges in $E_+$.
We set $\sigma(e)=1$ for each $e\in E_+$, and set $\sigma(f)=-1$ for each $f\in E_-$.
For a simplex $s$ of $T$, let $\Gamma_s$ denote the stabilizer of $s$ in $\Gamma$.
We pick two distinct vertices $v_0, v\in V(T)$.
Let $e_1,\ldots, e_n$ be the shortest sequence of oriented edges of $T$ such that the origin of $e_1$ is $v_0$; the terminal of $e_n$ is $v$; and for any $i=1,\ldots, n-1$, the terminal of $e_i$ and the origin of $e_{i+1}$ are equal.
We define $M$ as the maximal number in the set $\{ 0\} \cup \{\, \sum_{i=1}^k\sigma(e_i)\mid k=1,\ldots, n\, \}$, and define $m$ as the minimal number in it.
The equality
\[[\Gamma_{v_0}: \Gamma_{v_0}\cap \Gamma_v]=d_0|p_0|^m|q_0|^M\]
then holds, where we denote by $d_0>0$ the greatest common divisor of $p$ and $q$, and set $p_0=p/d_0$ and $q_0=q/d_0$.
For a group $G$ and a subgroup $H$ of $G$, we set
\[{\rm Comm}_G(H)=\{\, g\in G\mid [H: gHg^{-1}\cap H]<\infty,\ [gHg^{-1}:gHg^{-1}\cap H]<\infty\,\}\]
and call it the {\it (relative) commensurator} of $H$ in $G$, which is a subgroup of $G$.
We say that an element $\gamma$ of $\Gamma$ is {\it elliptic} if it fixes a vertex of $T$.
This condition is equivalent to the equality ${\rm Comm}_{\Gamma}(\langle \gamma \rangle)=\Gamma$ (see \cite[Lemma 2.1]{levitt}).
It follows that ellipticity of elements of $\Gamma$ is independent of the presentation (\ref{pre}).
We say that a subgroup $E$ of $\Gamma$ is {\it elliptic} if it fixes a vertex of $T$.
This condition is equivalent to that there exists an elliptic element of $\Gamma$ generating $E$.
It follows that ellipticity of subgroups of $\Gamma$ is also independent of the presentation (\ref{pre}).
We define the {\it modular homomorphism} $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$ by sending $a$ to $1$ and $t$ to $|q/p|$.
This is indeed independent of the presentation (\ref{pre}) as shown below.
Any two non-trivial elliptic subgroups $E$, $F$ of $\Gamma$ are commensurable, that is, the intersection $E\cap F$ is of finite index in both $E$ and $F$.
Let $E$ be a non-trivial elliptic subgroup of $\Gamma$, and pick a non-neutral element $x$ of $E$.
For each $\gamma \in \Gamma$, since $\gamma E\gamma^{-1}$ and $E$ are commensurable, there exist non-zero integers $n$, $m$ with $\gamma x^n\gamma^{-1}=x^m$.
We define a map $f\colon \Gamma \to \mathbb{R}_+^{\times}$ by $f(\gamma)=|m/n|$.
This is a well-defined homomorphism, and is independent of the choice of $E$ and $x$.
Since we have $f(a)=1$ and $f(t)=|q/p|$ by definition, the equality $\mathfrak{m} =f$ holds.
The modular homomorphism $\mathfrak{m}$ is thus defined independently of the presentation (\ref{pre}).
\section{Discrete measured groupoids}\label{sec-dmg}
\subsection{Terminology}
We recommend the reader to consult \cite{kechris} and \cite[Chapter XIII, Section 3]{take3} for basic knowledge of standard Borel spaces and discrete measured groupoids, respectively.
Unless otherwise stated, all sets and maps that appear in this paper are assumed to be Borel, and relations among Borel sets and maps are understood to hold up to sets of measure zero.
We refer to a standard Borel space with a finite positive measure as a {\it standard finite measure space}.
When the measure is a probability one, we refer to it as a {\it standard probability space}.
Let $(X, \mu)$ be a standard finite measure space.
Let $\mathcal{G}$ be a discrete measured groupoid on $(X, \mu)$.
Let $r, s\colon \mathcal{G}\rightarrow X$ denote the range and source maps, respectively.
For a Borel subset $A\subset X$ of positive measure, we define
\[(\mathcal{G})_A=\{\, g\in \mathcal{G}\mid r(g), s(g)\in A\, \}\]
and call it the restriction of $\mathcal{G}$ to $A$.
We denote by $\mathcal{G}A$ the saturation
\[\mathcal{G}A=\{\, r(g)\in X\mid g\in \mathcal{G},\ s(g)\in A\, \},\]
which is a Borel subset of $X$.
We say that $\mathcal{G}$ is {\it finite} if for a.e.\ $x\in X$, the set $r^{-1}(x)$ consists of at most finitely many points.
We say that $\mathcal{G}$ is of {\it infinite type} if for a.e.\ $x\in X$, the set $r^{-1}(x)$ consists of infinitely many points.
Let $\Gamma$ be a discrete group.
We say that a Borel action of $\Gamma$ on $(X, \mu)$ is {\it non-singular} if the action preserves the class of the measure $\mu$.
For a non-singular action $\Gamma \c (X, \mu)$, the associated discrete measured groupoid is denoted by $\Gamma \ltimes (X, \mu)$.
If $\mu$ is not specified, it is also denoted by $\Gamma \ltimes X$.
The range and source maps of $\Gamma \ltimes (X, \mu)$ are defined by $r(\gamma, x)=\gamma x$ and $s(\gamma, x)=x$, respectively, for $\gamma \in \Gamma$ and $x\in X$.
The product operation is defined by $(\gamma_1, \gamma_2x)(\gamma_2, x)=(\gamma_1\gamma_2, x)$ for $\gamma_1, \gamma_2\in \Gamma$ and $x\in X$.
Let $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ be ergodic f.f.m.p.\ actions of discrete groups on standard finite measure spaces, and let $\mathcal{G}$ and $\mathcal{H}$ be the associated groupoids, respectively.
The two actions are WOE if and only if there are Borel subsets $A\subset X$ and $B\subset Y$ of positive measure such that $(\mathcal{G})_A$ and $(\mathcal{H})_B$ are isomorphic.
We recall the ergodic decomposition for a discrete measured groupoid.
\begin{thm}[\ci{Theorem 6.1}{hahn}]
Let $(X, \mu)$ be a standard finite measure space.
Let $\mathcal{G}$ be a groupoid on $X$ admitting the structure of a discrete measured groupoid on $(X, \mu)$.
Then there exist a standard finite measure space $(Z, \xi)$ and a Borel map $\pi \colon X\to Z$ such that $\pi_*\mu=\xi$; and for a.e.\ $z\in Z$, $\mathcal{G}$ admits the structure of an ergodic discrete measured groupoid on $(X, \mu_z)$, where $\mu_z$ is the probability measure on $X$ obtained through the disintegration $\mu =\int_Z\mu_zd\xi(z)$ with respect to $\pi$.
Moreover, if a standard finite measure space $(W, \omega)$ and a Borel map $\theta\colon X\to W$ satisfy these properties in place of $(Z, \xi)$ and $\pi$, respectively, then there exists a Borel isomorphism $f$ from a conull Borel subset of $Z$ onto a conull Borel subset of $W$ such that $\theta(x)=f\circ \pi(x)$ for a.e.\ $x\in X$.
\end{thm}
We call the map $\pi \colon (X, \mu)\to (Z, \xi)$ the {\it ergodic decomposition} for $\mathcal{G}$.
For $z\in Z$, we put $X_z=\pi^{-1}(z)$ and denote by $\mathcal{G}_z$ the ergodic discrete measured groupoid $\mathcal{G}$ on $(X, \mu_z)$.
Since $\mu_z$ is supported on $X_z$, we often identify $\mathcal{G}_z$ with a discrete measured groupoid on $(X_z, \mu_z)$.
We introduce an invariant map for an action of a groupoid.
If the unit space of the groupoid consists of a single point, then it is a fixed point of the action of the group.
\begin{defn}\label{defn-inv-map}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$, $\Gamma$ a discrete group, and $S$ a standard Borel space.
Suppose that we have a Borel action of $\Gamma$ on $S$ and a Borel homomorphism $\rho \colon \mathcal{G}\rightarrow \Gamma$.
A Borel map $\varphi \colon X\rightarrow S$ is said to be {\it $(\mathcal{G}, \rho)$-invariant} if we have the equality
\[\rho(g)\varphi(s(g))=\varphi(r(g))\]
for a.e. $g\in \mathcal{G}$.
More generally, if $A$ is a Borel subset of $X$ with positive measure and if a Borel map $\varphi \colon A\rightarrow S$ satisfies the above equality for a.e.\ $g\in (\mathcal{G})_A$, then we also say that $\varphi$ is {\it $(\mathcal{G}, \rho)$-invariant}.
\end{defn}
The following extendability of invariant maps is in general use.
\begin{lem}[\ci{Lemma 2.3}{kida-ama}]\label{lem-inv-ext}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Suppose that we have a Borel action of a discrete group $\Gamma$ on a standard Borel space $S$ and a Borel homomorphism $\rho \colon \mathcal{G}\rightarrow \Gamma$.
If $A$ is a Borel subset of $X$ with positive measure and if $\varphi \colon A\rightarrow S$ is a $(\mathcal{G}, \rho)$-invariant Borel map, then $\varphi$ extends to a $(\mathcal{G}, \rho)$-invariant Borel map from $\mathcal{G}A$ into $S$.
\end{lem}
We refer to \cite{ar-book} for amenable groupoids.
A consequence of amenability in terms of the fixed point property is the following:
\begin{prop}[\ci{Theorem 4.2.7}{ar-book}]\label{prop-ame-basic}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$, $\Gamma$ a discrete group, and $\rho \colon \mathcal{G}\rightarrow \Gamma$ a Borel homomorphism.
Suppose that $\Gamma$ acts on a compact Polish space $K$ continuously.
Let $M(K)$ denote the space of probability measures on $K$, on which $\Gamma$ naturally acts.
If $\mathcal{G}$ is amenable, then there exists a $(\mathcal{G}, \rho)$-invariant Borel map from $X$ into $M(K)$.
\end{prop}
\subsection{Index for groupoids}
The index of a subrelation in a discrete measured equivalence relation is introduced by Feldman, Sutherland and Zimmer \cite{fsz}.
Their definition is directly generalized to that for discrete measured groupoids.
We define index for groupoids and present its basic properties.
Let $(X, \mu)$ be a standard finite measure space and $\mathcal{G}$ a discrete measured groupoid on $(X, \mu)$.
Let $r, s\colon \mathcal{G}\to X$ denote the range and source maps of $\mathcal{G}$, respectively.
Let $\mathcal{H}$ be a subgroupoid of $\mathcal{G}$.
For each $x\in X$, we define an equivalence relation on $s^{-1}(x)$ so that two elements $g, h\in s^{-1}(x)$ are equivalent if and only if $gh^{-1}\in \mathcal{H}$.
For a subset $E$ of $s^{-1}(x)$, let $E/\mathcal{H}$ denote the set of equivalence classes with respect to this equivalence relation on $E$.
We define the {\it index} of $\mathcal{H}$ in $\mathcal{G}$ at $x$, denoted by $[\mathcal{G}: \mathcal{H}]_x$, as the cardinality $|s^{-1}(x)/\mathcal{H}|$.
\begin{lem}\label{lem-index}
In the above notation, we put $I(x)=[\mathcal{G}:\mathcal{H}]_x$ for $x\in X$.
Then the following assertions hold:
\begin{enumerate}
\item The function $I\colon X\to \mathbb{Z}_{>0}\cup \{ \infty \}$ is Borel, and we have $I(s(g))=I(r(g))$ for any $g\in \mathcal{G}$.
In particular, if $\mathcal{G}$ is ergodic, then $I$ is essentially constant.
\item For any Borel subset $A$ of $X$ with positive measure and any $x\in A$, we have the inequality $[(\mathcal{G})_A: (\mathcal{H})_A]_x\leq I(x)$.
If $\mathcal{H}$ is ergodic, then the equality holds.
\item If we have a non-singular action of a discrete group $\Gamma$ on $(X, \mu)$ and a subgroup $\Lambda$ of $\Gamma$ such that $\mathcal{G}=\Gamma \ltimes (X, \mu)$ and $\mathcal{H}=\Lambda \ltimes (X, \mu)$, then $I(x)=[\Gamma: \Lambda]$ for any $x\in X$.
\end{enumerate}
\end{lem}
\begin{proof}
We prove assertion (i), following the proof of \cite[Lemma 1.1 (a)]{fsz}.
Taking Borel sections of the source map $s$, we find a countable set $N$ and a Borel map $\eta_n\colon D_n\to \mathcal{G}$ indexed by $n\in N$, where $D_n$ is a Borel subset of $X$ with positive measure, such that for any $n\in N$ and any $x\in D_n$, we have $s\circ \eta_n(x)=x$; and for any $g\in \mathcal{G}$, there exists a unique $n\in N$ with $s(g)\in D_n$ and $\eta_n(s(g))=g$.
Similarly, we find a countable set $M$ and a Borel map $\zeta_m\colon D_m\to \mathcal{H}$ indexed by $m\in M$, where $D_m$ is a Borel subset of $X$ with positive measure, such that for any $m\in M$ and any $x\in D_m$, we have $s\circ \zeta_m(x)=x$; and for any $g\in \mathcal{H}$, there exists a unique $m\in M$ with $s(g)\in D_m$ and $\zeta_m(s(g))=g$.
For $n_1,\ldots, n_k\in N$ and $m\in M$, we define a Borel subset $E(n_1,\ldots, n_k; m)$ of $X$ as the set of all points $x$ of $X$ such that for any $i=1,\ldots, k$, we have $x\in D_{n_i}$; and for any distinct $i, j=1,\ldots, k$ with $r\circ \eta_{n_i}(x)\in D_m$, the two elements $\zeta_m(r\circ \eta_{n_i}(x))\eta_{n_i}(x)$ and $\eta_{n_j}(x)$ are distinct.
For any $k\in \mathbb{Z}_{>0}$, the equality
\[\{\, x\in X\mid I(x)\geq k\,\}=\bigcup_{(n_1,\ldots, n_k)\in N^k}\bigcap_{m\in M}E(n_1,\ldots, n_k; m)\]
holds.
The function $I$ is thus Borel.
For any $g\in \mathcal{G}$, putting $x=s(g)$ and $y=r(g)$, we have the bijection from $s^{-1}(x)$ onto $s^{-1}(y)$ sending each element $h$ of $s^{-1}(x)$ to $hg^{-1}$.
It induces a bijection from $s^{-1}(x)/\mathcal{H}$ onto $s^{-1}(y)/\mathcal{H}$.
The equality $I(x)=I(y)$ follows.
Assertion (i) is proved.
Let $A$ be a Borel subset of $X$ with positive measure.
Let $s_A\colon (\mathcal{G})_A\to A$ denote the source map of $(\mathcal{G})_A$.
The inequality in assertion (ii) holds because for any $x\in A$, the inclusion of $s_A^{-1}(x)$ into $s^{-1}(x)$ induces an injective map from $s_A^{-1}(x)/(\mathcal{H})_A$ into $s^{-1}(x)/\mathcal{H}$.
Assume that $\mathcal{H}$ is ergodic.
For $x\in X$, let $e_x\in \mathcal{G}$ denote the unit element at $x$.
Ergodicity of $\mathcal{H}$ implies that there exists a Borel map $f\colon X\to \mathcal{H}$ with $f(x)=e_x$ for any $x\in A$ and with $s\circ f(y)=y$ and $r\circ f(y)\in A$ for any $y\in X$.
For any $x\in A$ and any $g\in s^{-1}(x)$, the two elements $g$ and $f(r(g))g$ lie in the same class in $s^{-1}(x)/\mathcal{H}$, and we have $f(r(g))g\in (\mathcal{G})_A$.
The map from $s_A^{-1}(x)/(\mathcal{H})_A$ into $s^{-1}(x)/\mathcal{H}$ is therefore surjective, and the equality $[(\mathcal{G})_A: (\mathcal{H})_A]_x=I(x)$ follows.
Assertion (ii) is proved.
Assertion (iii) follows by definition.
\end{proof}
\begin{lem}\label{lem-index-formula}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $r, s\colon \mathcal{G}\to X$ denote the range and source maps of $\mathcal{G}$, respectively.
Let $\mathcal{H}$, $\mathcal{K}$ and $\mathcal{L}$ be subgroupoids of $\mathcal{G}$.
Then the following assertions hold:
\begin{enumerate}
\item For any $x\in X$, we have $[\mathcal{G}:\mathcal{H}\cap \mathcal{K}]_x\leq [\mathcal{G}:\mathcal{H}]_x[\mathcal{G}:\mathcal{K}]_x$.
\item If $\mathcal{H}<\mathcal{K}$, then for any $x\in X$, we have $[\mathcal{K}\cap \mathcal{L}: \mathcal{H}\cap \mathcal{L}]_x\leq [\mathcal{K}: \mathcal{H}]_x$.
\item If $\mathcal{H}<\mathcal{K}$, then for any $x\in X$, we have the equality
\[[\mathcal{G}: \mathcal{H}]_x=\sum_{g\in E}[\mathcal{K}:\mathcal{H}]_{r(g)},\]
where $E$ is a set of representatives of all classes in $s^{-1}(x)/\mathcal{K}$.
Moreover, if any real-valued Borel function on $X$ that is $\mathcal{K}$-invariant is $\mathcal{G}$-invariant, then for a.e.\ $x\in X$, we have the equality
\[[\mathcal{G}:\mathcal{H}]_x=[\mathcal{G}:\mathcal{K}]_x[\mathcal{K}:\mathcal{H}]_x.\]
\end{enumerate}
\end{lem}
\begin{proof}
For $z\in X$, we set $\mathcal{G}_z=s^{-1}(z)$.
We prove assertion (i).
Fix $x\in X$.
Let
\[p\colon \mathcal{G}_x/(\mathcal{H}\cap \mathcal{K})\to \mathcal{G}_x/\mathcal{H},\quad q\colon \mathcal{G}_x/(\mathcal{H}\cap \mathcal{K})\to \mathcal{G}_x/\mathcal{K}\]
be the natural maps.
For any $g, h\in \mathcal{G}_x$, if $p(g)=p(h)$ and $q(g)=q(h)$, then $gh^{-1}\in \mathcal{H}\cap \mathcal{K}$.
It follows that the map from the set $\mathcal{G}_x/(\mathcal{H}\cap \mathcal{K})$ into the product $\mathcal{G}_x/\mathcal{H}\times \mathcal{G}_x/\mathcal{K}$ defined by $p$ and $q$ is injective.
Assertion (i) is proved.
Assertion (ii) holds because for any $x\in X$, the inclusion of $\mathcal{G}_x\cap \mathcal{K}\cap \mathcal{L}$ into $\mathcal{G}_x\cap \mathcal{K}$ induces an injective map from $(\mathcal{G}_x\cap \mathcal{K}\cap \mathcal{L})/(\mathcal{H}\cap \mathcal{L})$ into $(\mathcal{G}_x\cap \mathcal{K})/\mathcal{H}$.
We prove assertion (iii).
For any $z\in X$, we have the natural map $q_z\colon \mathcal{G}_z\to \mathcal{G}_z/\mathcal{K}$.
Fix $x\in X$.
Let $E$ be a set of representatives of all classes in $\mathcal{G}_x/\mathcal{K}$.
Pick $g\in E$ and put $y=r(g)$.
Let $e_y\in \mathcal{G}$ denote the unit element at $y$.
The map from $q_x^{-1}(q_x(g))$ onto $q_y^{-1}(q_y(e_y))$ sending each element $h$ of $q_x^{-1}(q_x(g))$ to $hg^{-1}$ is bijective.
This map induces a bijection from $q_x^{-1}(q_x(g))/\mathcal{H}$ onto $q_y^{-1}(q_y(e_y))/\mathcal{H}$.
We thus have the equality $|q_x^{-1}(q_x(g))/\mathcal{H}|=|q_y^{-1}(q_y(e_y))/\mathcal{H}|=[\mathcal{K}:\mathcal{H}]_y$.
The equality
\[[\mathcal{G}: \mathcal{H}]_x=|\mathcal{G}_x/\mathcal{H}|=\sum_{g\in E}|q_x^{-1}(q_x(g))/\mathcal{H}|=\sum_{g\in E}[\mathcal{K}:\mathcal{H}]_{r(g)}\]
is obtained.
The former assertion in assertion (iii) is proved.
The function on $X$ assigning $[\mathcal{K}: \mathcal{H}]_z$ to each $z\in X$ is $\mathcal{K}$-invariant by Lemma \ref{lem-index} (i).
If the ergodic decompositions for $\mathcal{G}$ and $\mathcal{K}$ are the same map, then this function is also $\mathcal{G}$-invariant.
For a.e.\ $x\in X$ and any $g\in E$, the equality $[\mathcal{K}: \mathcal{H}]_{r(g)}=[\mathcal{K}: \mathcal{H}]_x$ then holds.
The latter assertion in assertion (iii) is proved.
\end{proof}
The following lemma is obtained from the definition of index.
\begin{lem}\label{lem-index-fi}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $\Gamma$ be a discrete group, and let $\Lambda$ be a subgroup of $\Gamma$.
If $\rho \colon \mathcal{G}\to \Gamma$ is a Borel homomorphism, then $\rho^{-1}(\Lambda)$ is a subgroupoid of $\mathcal{G}$, and for any $x\in X$, we have $[\mathcal{G}:\rho^{-1}(\Lambda)]_x\leq [\Gamma :\Lambda]$.
\end{lem}
Let $\Gamma$ be a discrete group, $\Lambda$ a subgroup of $\Gamma$ of finite index, and $S$ a space on which $\Gamma$ acts.
Choose a family $\{ \gamma_1,\ldots, \gamma_N\}$ of representatives for all right cosets of $\Lambda$ in $\Gamma$ with $N=[\Gamma: \Lambda]$.
For any fixed point $x_0$ in $S$ for the action of $\Lambda$, the subset $\{ \gamma_1^{-1}x_0,\ldots, \gamma_N^{-1}x_0\}$ of $S$ is an orbit for the action of $\Gamma$ and is thus fixed by $\Gamma$.
The following lemma is an analogue for groupoids.
\begin{lem}\label{lem-finite-index}
In the notation in the second paragraph of this subsection, let $\Gamma$ be a discrete group, $\rho \colon \mathcal{G}\to \Gamma$ a Borel homomorphism, and $S$ a standard Borel space on which $\Gamma$ acts by Borel automorphisms.
We denote by $\mathcal{F}(S)$ the Borel space of all non-empty finite subsets of $S$, on which $\Gamma$ naturally acts.
Suppose that the function $I(x)=[\mathcal{G}:\mathcal{H}]_x$ on $X$ is essentially constant, and its essential value, denoted by $N$, is finite.
Let $\phi_1,\ldots, \phi_N$ be Borel maps from $X$ into $\mathcal{G}$ such that for a.e.\ $x\in X$, $\{ \phi_1(x),\ldots, \phi_N(x)\}$ is a set of representatives of all classes in $s^{-1}(x)/\mathcal{H}$.
Then for any $(\mathcal{H}, \rho)$-invariant Borel map $\psi \colon X\to S$, the Borel map $\Psi \colon X\to \mathcal{F}(S)$ defined by
\[\Psi(x)=\{\, \rho(\phi_1(x))^{-1}\psi(r\circ \phi_1(x)),\ldots, \rho(\phi_N(x))^{-1}\psi(r\circ \phi_N(x))\,\}\]
for $x\in X$ is $(\mathcal{G}, \rho)$-invariant.
\end{lem}
\begin{proof}
We denote by $\mathfrak{S}(N)$ the symmetric group on the set $\{ 1,\ldots, N\}$.
For a.e.\ $g\in \mathcal{G}$ with $x=s(g)$ and $y=r(g)$, the set $\{ \phi_1(x)g^{-1},\ldots, \phi_N(x)g^{-1}\}$ is then a set of representatives of all classes in $s^{-1}(y)/\mathcal{H}$.
There thus exists a Borel homomorphism $\alpha \colon \mathcal{G}\to \mathfrak{S}(N)$ such that for a.e.\ $g\in \mathcal{G}$ with $x=s(g)$ and $y=r(g)$ and for any $i\in \{ 1,\ldots, N\}$, we have $\phi_i(x)g^{-1}\phi_{\alpha(g)(i)}(y)^{-1}\in \mathcal{H}$.
Putting $j=\alpha(g)(i)$, we have
\begin{align*}
\rho(g)\rho(\phi_i(x))^{-1}\psi(r\circ \phi_i(x))&=\rho(\phi_j(y))^{-1}\rho(\phi_j(y)g\phi_i(x)^{-1})\psi(r\circ \phi_i(x))\\
&=\rho(\phi_j(y))^{-1}\psi(r\circ \phi_j(y)),
\end{align*}
where the last equality holds because $\psi$ is $(\mathcal{H}, \rho)$-invariant.
The lemma follows.
\end{proof}
In Lemma \ref{lem-finite-index}, we note that for a.e.\ $x\in X$, exactly one of $\phi_1(x),\ldots, \phi_N(x)$, say $\phi_1(x)$, belongs to $\mathcal{H}$, and we then have $\rho(\phi_1(x))^{-1}\psi(r\circ \phi_1(x))=\psi(x)$ because $\psi$ is $(\mathcal{H}, \rho)$-invariant.
It follows that $\Psi(x)$ contains $\psi(x)$ for a.e.\ $x\in X$.
\begin{lem}\label{lem-erg-dec}
In the notation in the second paragraph of this subsection, suppose that $\mathcal{G}$ is measure-preserving.
We also suppose that the function $I(x)=[\mathcal{G}: \mathcal{H}]_x$ on $X$ is essentially constant, and its essential value, denoted by $N$, is finite.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ and $\theta \colon (X, \mu)\to (W, \omega)$ be the ergodic decompositions for $\mathcal{G}$ and $\mathcal{H}$, respectively.
Let $\sigma \colon (W, \omega)\to (Z, \xi)$ be the canonical Borel map such that $\pi =\sigma \circ \theta$, i.e., the following diagram commutes:
\[\xymatrix{
& \ar[dl]_{\theta} (X, \mu) \ar[dr]^{\pi} & \\
(W, \omega) \ar[rr]^{\sigma} & & (Z, \xi) \\
}\]
Let $\omega =\int_{Z}\omega_z d\xi(z)$ be the disintegration with respect to $\sigma$.
Then for a.e.\ $z\in Z$, the set $\sigma^{-1}(z)$ consists of at most $N$ points up to $\omega_z$-null sets.
\end{lem}
\begin{proof}
To prove the lemma, it is enough to show that for a.e.\ $z\in Z$, there exists no Borel subset of $\sigma^{-1}(z)$ whose measure with respect to $\omega_z$ is positive and less than $1/N$.
We assume the contrary, and will deduce a contradiction.
Thanks to general description of a measurable decomposition of a measure space in \cite[\S 4, No.1]{rohlin}, there exist a countable set $M$, a Borel partition $Z=\bigsqcup_{m\in M}Z_m$ and a standard finite measure space $(E_m, \eta_m)$ indexed by each $m\in M$ satisfying the following:
For any $m\in M$, there exists a Borel isomorphism $f_m$ from a conull Borel subset of $E_m\times Z_m$ onto a conull Borel subset of $\sigma^{-1}(Z_m)$ such that for a.e.\ $y\in E_m$ and a.e.\ $z\in Z_m$, the equality $\sigma \circ f_m(y, z)=z$ holds; and the measure $(f_m)_*(\eta_m\times \xi_m)$ and the restriction of $\omega$ to $\sigma^{-1}(Z_m)$ are equivalent, where $\xi_m$ is the restriction of $\xi$ to $Z_m$.
This application of the result in \cite{rohlin} owes to the proof of \cite[Proposition 2.21]{fmw}.
Our assumption in the last paragraph implies that there exists a Borel subset $A$ of $Z$ with $\xi(A)>0$ and a Borel subset $B$ of $\sigma^{-1}(A)$ with $0<\omega_z(B\cap \sigma^{-1}(z))<1/N$ for a.e.\ $z\in A$.
We choose Borel maps $\phi_1,\ldots, \phi_N$ from $X$ into $\mathcal{G}$ such that for a.e.\ $x\in X$, $\{ \phi_1(x),\ldots, \phi_N(x)\}$ is a set of representatives of all classes in $s^{-1}(x)/\mathcal{H}$.
We put
\[Y=\theta^{-1}(B)\quad \textrm{and}\quad D=\bigcup_{i=1}^N (r\circ \phi_i)^{-1}(Y).\]
We claim that $D$ is $\mathcal{G}$-invariant.
For a.e.\ $g\in \mathcal{G}$ with $x=s(g)\in D$ and $y=r(g)$, there exists $i$ with $r\circ \phi_i(x)\in Y$, and there exists $j$ with $\phi_j(y)g\phi_i(x)^{-1}\in \mathcal{H}$.
Since $Y$ is $\mathcal{H}$-invariant, we have $r\circ \phi_j(y)\in Y$.
We thus have $y\in D$.
The claim is proved.
The claim and the inclusion $B\subset \sigma^{-1}(A)$ imply the inclusion $D\subset \pi^{-1}(A)$.
We show the converse inclusion.
Let $\mu =\int_Z\mu_z d\xi(z)$ be the disintegration with respect to $\pi$.
We have $\omega =\theta_*\mu =\int_Z\theta_*\mu_z d\xi(z)$.
By uniqueness of disintegration, we have $\omega_z=\theta_*\mu_z$ for a.e.\ $z\in Z$.
The condition $\omega_z(B\cap \sigma^{-1}(z))>0$ for a.e.\ $z\in A$ implies that $\mu_z(Y\cap \pi^{-1}(z))>0$ for a.e.\ $z\in A$.
Since $D$ is a $\mathcal{G}$-invariant Borel subset of $X$ containing $Y$, we have $\pi^{-1}(A)\subset D$.
We thus obtained the equality $D=\pi^{-1}(A)$.
By the definition of $D$, we have
\[\mu(D)\leq N\mu(Y)=N\omega(B)=N\int_A\omega_z(B\cap \sigma^{-1}(z))\, d\xi(z)<\xi(A)=\mu(D).\]
This is a contradiction.
\end{proof}
In Lemma \ref{lem-erg-dec}, thanks to its conclusion, replacing $W$ by its conull Borel subset, we can assume that for a.e.\ $z\in Z$, any point of $\sigma^{-1}(z)$ has positive measure with respect to $\omega_z$, and $\sigma^{-1}(z)$ consists of at most $N$ points.
We end this subsection with the following lemma on disintegration of a measure, which will be applied in the setting of Lemma \ref{lem-erg-dec}.
\begin{lem}\label{lem-rest}
Let $(X, \mu)$, $(Z, \xi)$ and $(W, \omega)$ be standard finite measure spaces.
Let $\pi \colon X\to Z$, $\theta \colon X\to W$ and $\sigma \colon W\to Z$ be Borel maps satisfying the equalities $\pi_*\mu=\xi$, $\theta_*\mu=\omega$ and $\pi =\sigma \circ \theta$.
Let $\mu =\int_Z\mu_zd\xi(z)$ and $\mu =\int_W\nu_wd\omega(w)$ be the disintegrations with respect to $\pi$ and $\theta$, respectively.
Suppose that for a.e.\ $z\in Z$, the set $\sigma^{-1}(z)$ is countable.
Then for a.e.\ $w\in W$, we have the equality
\[\mu_{\sigma(w)}|_{\theta^{-1}(w)}=\mu_{\sigma(w)}(\theta^{-1}(w))\nu_w.\]
\end{lem}
\begin{proof}
For $z\in Z$, we put $X_z=\pi^{-1}(z)$ and $W_z=\sigma^{-1}(z)$.
Similarly, for $w\in W$, we put $X_w=\theta^{-1}(w)$.
For a.e.\ $z\in Z$, we have the partition $X_z=\bigsqcup_{w\in W_z}X_w$ into countably many Borel subsets.
For any Borel subset $B$ of $W$, we have
\begin{align*}
\omega(B)&=\mu(\theta^{-1}(B))=\int_Z\mu_z(\theta^{-1}(B)\cap X_z)\, d\xi(z)=\int_Z\sum_{w\in B\cap W_z}\mu_z(X_w)\, d\xi(z).
\end{align*}
For any Borel subset $A$ of $X$, we thus have
\begin{align*}
\mu(A)&=\int_Z\mu_z(A\cap X_z)\, d\xi(z)=\int_Z\sum_{w\in W_z}\mu_z(A\cap X_w)\, d\xi(z)\\
&=\int_W\frac{\mu_{\sigma(w)}(A\cap X_w)}{\mu_{\sigma(w)}(X_w)}\, d\omega(w).
\end{align*}
By uniqueness of disintegration, we obtain the desired equality.
\end{proof}
\subsection{Local index}
Throughout this subsection, we fix a standard finite measure space $(X, \mu)$, a discrete measured groupoid $\mathcal{G}$ on $(X, \mu)$ and a subgroupoid $\mathcal{H}$ of $\mathcal{G}$.
As proved in Lemma \ref{lem-index} (ii), if $A$ is a Borel subset of $X$ with positive measure, the inequality $[(\mathcal{G})_A:(\mathcal{H})_A]_x\leq [\mathcal{G}:\mathcal{H}]_x$ holds for any $x\in A$ though the equality does not hold in general.
Under a certain assumption, for $x\in X$, we define the local index of $\mathcal{H}$ in $\mathcal{G}$ at $x$, denoted by $[[\mathcal{G}: \mathcal{H}]]_x$, which necessarily satisfies the equality $[[(\mathcal{G})_A:(\mathcal{H})_A]]_x=[[\mathcal{G}:\mathcal{H}]]_x$ for any Borel subset $A$ of $X$ with positive measure and a.e.\ $x\in A$.
This local index can be seen as an anlogue of the local index for subfactors introduced by Jones in \cite[\S 2.2]{jones}.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ and $\theta \colon (X, \mu)\to (W, \omega)$ be the ergodic decompositions for $\mathcal{G}$ and $\mathcal{H}$, respectively.
Let $\sigma \colon (W, \omega)\to (Z, \xi)$ be the canonical Borel map such that $\pi =\sigma \circ \theta$, i.e., the following diagram commutes:
\[\xymatrix{
& \ar[dl]_{\theta} (X, \mu) \ar[dr]^{\pi} & \\
(W, \omega) \ar[rr]^{\sigma} & & (Z, \xi) \\
}\]
We suppose that for any $z\in Z$, the set $\sigma^{-1}(z)$ is countable.
Under this assumption, we define the local index of $\mathcal{H}$ in $\mathcal{G}$ at $x\in X$.
Choose a countable set $N$ and a Borel partition $W=\bigsqcup_{n\in N}W_n$ such that for any $n\in N$, we have $\omega(W_n)>0$, and the map $\sigma$ is injective on $W_n$.
Set $Y_n=\theta^{-1}(W_n)$ for $n\in N$.
We define a Borel function $J\colon X\to \mathbb{Z}_{>0}\cup \{ \infty \}$ so that for $n\in N$ and $x\in Y_n$, we have $J(x)=[(\mathcal{G})_{Y_n}: (\mathcal{H})_{Y_n}]_x$.
The following lemma implies that this function does not depend on the choice of the partition $W=\bigsqcup_{n\in N}W_n$.
\begin{lem}\label{lem-li-wd}
In the above notation, fix $n\in N$.
If $A$ is a Borel subset of $Y_n$ with $\mu(A)>0$, then for a.e.\ $x\in A$, we have $[(\mathcal{G})_A:(\mathcal{H})_A]_x=[(\mathcal{G})_{Y_n}:(\mathcal{H})_{Y_n}]_x$.
\end{lem}
\begin{proof}
We put $Y=Y_n$.
Let $s_Y\colon (\mathcal{G})_Y\to Y$ and $s_A\colon (\mathcal{G})_A\to A$ be the source maps.
For any $x\in A$, the inclusion of $s_A^{-1}(x)$ into $s_Y^{-1}(x)$ induces the injective map $\imath$ from $s_A^{-1}(x)/\mathcal{H}$ into $s_Y^{-1}(x)/\mathcal{H}$.
To prove the lemma, it suffices to show that $\imath$ is surjective for a.e.\ $x\in A$.
Since $\sigma$ is injective on $W_n$, the equality $(\mathcal{G})_YA=\mathcal{H}A$ holds.
For a.e.\ $x\in A$ and any $g\in s_Y^{-1}(x)$, there thus exists $h\in \mathcal{H}$ with $s(h)\in A$ and $r(h)=r(g)$.
The product $h^{-1}g$ belongs to $s_A^{-1}(x)$, and we have $g(h^{-1}g)^{-1}=h\in \mathcal{H}$.
The map $\imath$ is therefore surjective.
\end{proof}
For $x\in X$, we denote $J(x)$ by $[[\mathcal{G}:\mathcal{H}]]_x$, and call this number the {\it local index} of $\mathcal{H}$ in $\mathcal{G}$ at $x$.
We present basic properties of local index.
\begin{lem}\label{lem-li-prod}
In the notation in the second paragraph of this subsection, let $\mathcal{K}$ be a subgroupoid of $\mathcal{G}$ with $\mathcal{H}<\mathcal{K}$.
Then for a.e.\ $x\in X$, the local indices $[[\mathcal{G}:\mathcal{K}]]_x$ and $[[\mathcal{K}:\mathcal{H}]]_x$ at $x$ are well-defined, and we have the equality
\[[[\mathcal{G}:\mathcal{H}]]_x=[[\mathcal{G}:\mathcal{K}]]_x[[\mathcal{K}:\mathcal{H}]]_x.\]
\end{lem}
\begin{proof}
Let $\Phi \colon (X, \mu)\to (V, \upsilon)$ be the ergodic decomposition for $\mathcal{K}$.
We have the canonical Borel map from $(W, \omega)$ into $(V, \upsilon)$ and that from $(V, \upsilon)$ into $(Z, \xi)$ because we have $\mathcal{H}<\mathcal{K}<\mathcal{G}$.
Since the composition of these two maps is equal to $\sigma$, the inverse image of any point under these two maps is countable.
The former assertion follows.
The desired equality follows from Lemma \ref{lem-index-formula} (iii) and Lemma \ref{lem-li-wd}.
\end{proof}
\begin{lem}\label{lem-li-res}
In the notation in the second paragraph of this subsection, if $A$ is a Borel subset of $X$ with $\mu(A)>0$, then for a.e.\ $x\in A$, we have the equality
\[[[(\mathcal{G})_A:(\mathcal{H})_A]]_x=[[\mathcal{G}:\mathcal{H}]]_x.\]
\end{lem}
\begin{proof}
Let $\pi_1\colon (A, \mu|_A)\to (Z_1, \xi_1)$ and $\theta_1\colon (A, \mu|_A)\to (W_1, \omega_1)$ be the ergodic decompositions for $(\mathcal{G})_A$ and $(\mathcal{H})_A$, respectively.
We have the canonical Borel map $\sigma_1\colon (W_1, \omega_1)\to (Z_1, \xi_1)$ such that $\pi_1=\sigma_1\circ \theta_1$, i.e., the following diagram commutes:
\[\xymatrix{
& \ar[dl]_{\theta_1} (A, \mu|_A) \ar[dr]^{\pi_1} & \\
(W_1, \omega_1) \ar[rr]^{\sigma_1} & & (Z_1, \xi_1) \\
}\]
We also have the canonical injective Borel map from $Z_1$ into $Z$, and identify $Z_1$ with the image, which is a Borel subset of $Z$.
In the same manner, we naturally identify $W_1$ with a Borel subset of $W$.
The map $\sigma_1$ is the restriction of $\sigma$ under these identifications.
Let $W_2$ be a Borel subset of $W_1$ such that $\omega_1(W_2)>0$, and $\sigma_1$ is injective on $W_2$.
Put $Y_1=\theta_1^{-1}(W_2)$.
By definition, we have $[[(\mathcal{G})_A:(\mathcal{H})_A]]_x=[(\mathcal{G})_{Y_1}:(\mathcal{H})_{Y_1}]_x$ for a.e.\ $x\in Y_1$.
We set $Y=\mathcal{H}Y_1=\theta^{-1}(W_2)$.
Since $\sigma$ is injective on $W_2$, we have the equality $[[\mathcal{G}:\mathcal{H}]]_x=[(\mathcal{G})_Y:(\mathcal{H})_Y]_x$ for a.e.\ $x\in Y$.
By Lemma \ref{lem-li-wd}, the desired equality holds for a.e.\ $x\in Y_1$.
\end{proof}
\begin{lem}\label{lem-li-group}
Let $\Gamma$ be a discrete group.
Let $\Lambda$ be a finite index, normal subgroup of $\Gamma$.
Suppose that we have a non-singular action of $\Gamma$ on a standard finite measure space $(X, \mu)$.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{H}=\Lambda \ltimes X$, and define the maps $\pi$, $\theta$ and $\sigma$ as in the second paragraph of this subsection.
Let $\omega =\int_Z\omega_zd\xi(z)$ be the disintegration with respect to $\sigma$.
We assume that for a.e.\ $z\in Z$, any point of $\sigma^{-1}(z)$ has positive measure with respect to $\omega_z$.
Then for a.e.\ $x\in X$, we have the equality
\[[[\mathcal{G}:\mathcal{H}]]_x=\frac{[\Gamma :\Lambda]}{|\sigma^{-1}(\pi(x))|}.\]
\end{lem}
\begin{proof}
We have the canonical non-singular action of $\Gamma$ on $(W, \omega)$ because $\Lambda$ is normal in $\Gamma$.
For $w\in W$, let $\Gamma_w$ denote the stabilizer of $w$ in $\Gamma$, which contains $\Lambda$.
The set $\sigma^{-1}(\sigma(w))$ is then identified with $\Gamma /\Gamma_w$.
Let $W_1$ be a Borel subset of $W$ such that $\omega(W_1)>0$; the map $\sigma$ is injective on $W_1$; and $\Gamma_w=\Gamma_{w'}$ for any $w, w'\in W_1$.
Set $Y=\theta^{-1}(W_1)$ and fix $w_0\in W_1$.
The set $Y$ is $\Gamma_{w_0}$-invariant, and the equality $(\mathcal{G})_Y=\Gamma_{w_0}\ltimes Y$ holds.
For a.e.\ $x\in Y$, we then have the equality
\[[[\mathcal{G}:\mathcal{H}]]_x=[(\mathcal{G})_Y:(\mathcal{H})_Y]_x=[\Gamma_{w_0}\ltimes Y: \Lambda \ltimes Y]_x=[\Gamma_{w_0}:\Lambda].\]
The lemma therefore follows.
\end{proof}
\subsection{Quasi-normal subgroupoids}
Normal subgroupoids are studied in \cite{kida-exama} and \cite{sauer-thom} as a generalization of normal subrelations introduced by Feldman, Sutherland and Zimmer \cite{fsz}.
In this subsection, we recall normal subgroupoids, and introduce quasi-normal subgroupoids, which are a slight generalization of normal ones.
We refer to \cite{aoi} and \cite{ay} for related works in the framework of von Neumann algebras.
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $r, s\colon \mathcal{G}\to X$ denote the range and source maps of $\mathcal{G}$, respectively.
We define $[[\mathcal{G}]]$ as the set of all Borel maps $\phi \colon D_{\phi}\to \mathcal{G}$ from a Borel subset $D_{\phi}$ of $X$ into $\mathcal{G}$ such that $s\circ \phi(x)=x$ for any $x\in D_{\phi}$; and the map $r\circ \phi \colon D_{\phi}\to X$ is injective.
For any $\phi \in [[\mathcal{G}]]$, we set $R_{\phi}=r\circ \phi(D_{\phi})$ and define a Borel map $U_{\phi}\colon (\mathcal{G})_{D_{\phi}}\to (\mathcal{G})_{R_{\phi}}$ by the formula $U_{\phi}(g)=\phi(r(g))g\phi(s(g))^{-1}$ for $g\in (\mathcal{G})_{D_{\phi}}$.
The map $U_{\phi}$ is an isomorphism of discrete measured groupoids.
Let $\mathcal{S}$ be a subgroupoid of $\mathcal{G}$.
For $\phi \in [[\mathcal{G}]]$, we set $\mathcal{S}^{\phi}=U_{\phi}((\mathcal{S})_{D_{\phi}})$.
We define two subsets ${\rm N}_{\mathcal{G}}(\mathcal{S})$, ${\rm QN}_{\mathcal{G}}(\mathcal{S})$ of $[[\mathcal{G}]]$ by
\begin{align*}
{\rm N}_{\mathcal{G}}(\mathcal{S})&=\{\, \phi \in [[\mathcal{G}]]\mid \mathcal{S}^{\phi}=(\mathcal{S})_{R_{\phi}}\,\},\\
{\rm QN}_{\mathcal{G}}(\mathcal{S})&=\{\, \phi \in [[\mathcal{G}]]\mid [(\mathcal{S})_{R_{\phi}}: (\mathcal{S})_{R_{\phi}}\cap \mathcal{S}^{\phi}]_x<\infty,\ [\mathcal{S}^{\phi}: (\mathcal{S})_{R_{\phi}}\cap \mathcal{S}^{\phi}]_x<\infty\\
& \hspace{85mm} \textrm{for a.e.\ }x\in R_{\phi}\,\}.
\end{align*}
\begin{defn}\label{defn-qn}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$ with $s\colon \mathcal{G}\to X$ the source map.
Let $\mathcal{S}$ be a subgroupoid of $\mathcal{G}$.
\begin{enumerate}
\item We say that $\mathcal{S}$ is {\it normal} in $\mathcal{G}$ if there exists a countable family $\{ \phi_n\}_n$ of elements of ${\rm N}_{\mathcal{G}}(\mathcal{S})$ such that for a.e.\ $g\in \mathcal{G}$, there exists $n$ with $s(g)\in D_{\phi_n}$ and $\phi_n(s(g))g^{-1}\in \mathcal{S}$.
\item We say that $\mathcal{S}$ is {\it quasi-normal} in $\mathcal{G}$ if the same condition as in (i) holds after replacing ${\rm N}_{\mathcal{G}}(\mathcal{S})$ with ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
\end{enumerate}
\end{defn}
While the above definition of normal subgroupoids is slightly different from those in \cite{kida-exama} and \cite{sauer-thom}, it can be checked that they are equivalent.
We introduce composition and inverse of elements of $[[\mathcal{G}]]$.
Pick $\phi, \psi \in [[\mathcal{G}]]$.
We define $\eta \in [[\mathcal{G}]]$ as follows.
Set $D_{\eta}=(r\circ \phi)^{-1}(D_{\psi}\cap R_{\phi})$ and define a Borel map $\eta \colon D_{\eta}\to \mathcal{G}$ by $\eta(x)=\psi(r\circ \phi(x))\phi(x)$ for $x\in D_{\eta}$.
The map $\eta$ then belongs to $[[\mathcal{G}]]$.
The isomorphism $U_{\eta}$ is equal to the restriction of $U_{\psi}\circ U_{\phi}$ to $(\mathcal{G})_{D_{\eta}}$.
We denote the map $\eta$ by $\psi \bullet \phi$ and call it the {\it composition} of $\phi$ and $\psi$.
For $\phi \in [[\mathcal{G}]]$, we define $\zeta \in [[\mathcal{G}]]$ as follows.
Set $D_{\zeta}=R_{\phi}$ and define a Borel map $\zeta \colon D_{\zeta}\to \mathcal{G}$ by $\zeta(x)=\phi((r\circ \phi)^{-1}(x))^{-1}$ for $x\in D_{\zeta}$.
The map $\zeta$ then belongs to $[[\mathcal{G}]]$.
We have the equality $U_{\zeta}=U_{\phi}^{-1}$.
Let us call the map $\zeta$ the {\it inverse} of $\phi$.
\begin{lem}\label{lem-qn}
Let $\mathcal{S}$ be a subgroupoid of $\mathcal{G}$.
Then the following assertions hold.
\begin{enumerate}
\item For any $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$ and any Borel subset $A$ of $D_{\phi}$, the restriction of $\phi$ to $A$, denoted by $\phi|_A$, belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
\item For any $\phi, \psi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$, we have $\psi \bullet \phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$.
\item For any $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$, the inverse of $\phi$ also belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
\end{enumerate}
\end{lem}
\begin{proof}
Let $r, s\colon \mathcal{G}\to X$ be the range and source maps of $\mathcal{G}$, respectively.
Assertion (i) follows from Lemma \ref{lem-index} (ii).
To prove assertion (iii), we pick $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$.
Let $\theta$ be the inverse of $\phi$.
For a.e.\ $x\in R_{\theta}$, we have
\[[(\mathcal{S})_{R_{\theta}}: (\mathcal{S})_{R_{\theta}}\cap \mathcal{S}^{\theta}]_x=[(\mathcal{S})_{D_{\phi}}: (\mathcal{S})_{D_{\phi}}\cap \mathcal{S}^{\theta}]_x=[\mathcal{S}^{\phi}: \mathcal{S}^{\phi}\cap (\mathcal{S})_{R_{\phi}}]_{r\circ \phi(x)},\]
where the second equality is obtained by applying $U_{\phi}$.
The right hand side is finite because $\phi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
In a similar way, we can show that $[\mathcal{S}^{\theta}: (\mathcal{S})_{R_{\theta}}\cap \mathcal{S}^{\theta}]_x$ is finite for a.e.\ $x\in R_{\theta}$.
Assertion (iii) is proved.
To prove assertion (ii), we pick $\phi, \psi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$ and put $\eta =\psi \bullet \phi$.
Let $\zeta$ be the inverse of $\psi$.
For any $x\in R_{\eta}$, let $E$ be a set of representatives of all classes in $(s^{-1}(x)\cap (\mathcal{S})_{R_{\eta}})/((\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}})$.
We have
\begin{align*}
&[(\mathcal{S})_{R_{\eta}}: (\mathcal{S})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_x\leq [(\mathcal{S})_{R_{\eta}}: (\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_x\\
= \, & \sum_{g\in E}[(\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}}: (\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_{r(g)},
\end{align*}
where the last equality holds by Lemma \ref{lem-index-formula} (iii).
For a.e.\ $x\in R_{\eta}$, the set $E$ is finite because $\psi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
We put $C=r\circ \zeta(R_{\eta})$.
For any $g\in E$, putting $y=r(g)$, we have
\begin{align*}
&[(\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}}: (\mathcal{S})_{R_{\eta}}\cap (\mathcal{S}^{\psi})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_y\\
= \, & [(\mathcal{S}^{\zeta})_C\cap (\mathcal{S})_C: (\mathcal{S}^{\zeta})_C\cap (\mathcal{S})_C\cap (\mathcal{S}^{\phi})_C]_{r\circ \zeta(y)}\\
\leq \, & [(\mathcal{S})_C: (\mathcal{S})_C\cap (\mathcal{S}^{\phi})_C]_{r\circ \zeta(y)}\leq [(\mathcal{S})_{R_{\phi}}: (\mathcal{S})_{R_{\phi}}\cap \mathcal{S}^{\phi}]_{r\circ \zeta(y)},
\end{align*}
where the first equality is obtained by applying $U_{\zeta}$.
The first and second inequalities follow from Lemma \ref{lem-index-formula} (ii) and Lemma \ref{lem-index} (ii), respectively.
The right hand side is finite because $\phi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
We therefore have $[(\mathcal{S})_{R_{\eta}}: (\mathcal{S})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_x<\infty$ for a.e.\ $x\in R_{\eta}$.
Let $\xi$ denote the inverse of $\eta$, which is the composition of the inverses of $\psi$ and of $\phi$.
For a.e.\ $x\in R_{\eta}$, applying $U_{\xi}$, we obtain the equality
\[[\mathcal{S}^{\eta}: (\mathcal{S})_{R_{\eta}}\cap \mathcal{S}^{\eta}]_x=[(\mathcal{S})_{R_{\xi}}: \mathcal{S}^{\xi}\cap (\mathcal{S})_{R_{\xi}}]_{r\circ \xi(x)}.\]
The right hand side is finite thanks to assertion (iii) and the argument in the last paragraph.
Assertion (ii) is proved.
\end{proof}
In the notation in Definition \ref{defn-qn}, $\mathcal{S}$ is normal in $\mathcal{G}$ if and only if there exists a countable family $\{ \phi_n\}_n$ of elements of ${\rm N}_{\mathcal{G}}(\mathcal{S})$ such that $\mathcal{G}=\bigcup_{n}\phi_n(D_{\phi_n})$ up to null sets.
This is because $[[\mathcal{S}]]$ is contained in ${\rm N}_{\mathcal{G}}(\mathcal{S})$ and the composition of two elements of ${\rm N}_{\mathcal{G}}(\mathcal{S})$ also belongs to ${\rm N}_{\mathcal{G}}(\mathcal{S})$.
A similar property holds for quasi-normality if ${\rm N}_{\mathcal{G}}(\mathcal{S})$ is replaced by ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
Let $\Gamma$ be a discrete group and $\Lambda$ a subgroup of $\Gamma$.
The set ${\rm QN}_{\Gamma}(\Lambda)$ introduced right before Definition \ref{defn-qn} is naturally identified with the subgroup ${\rm Comm}_{\Gamma}(\Lambda)$ of $\Gamma$ introduced in Section \ref{sec-bs}.
It follows that $\Lambda$ is quasi-normal in $\Gamma$ in the sense of Definition \ref{defn-qn} if and only if the equality ${\rm Comm}_{\Gamma}(\Lambda)=\Gamma$ holds.
\begin{lem}\label{lem-qn-group}
Let $\Gamma$ be a discrete group and $\Lambda$ a subgroup of $\Gamma$.
Let $\Gamma \c (X, \mu)$ be a non-singular action.
If $\Lambda$ is quasi-normal in $\Gamma$, then $\Lambda \ltimes (X, \mu)$ is quasi-normal in $\Gamma \ltimes (X, \mu)$.
\end{lem}
\begin{proof}
Put $\mathcal{G}=\Gamma \ltimes (X, \mu)$ and $\mathcal{H}=\Lambda \ltimes (X, \mu)$.
For any $\gamma \in \Gamma$, the map from $X$ into $\mathcal{G}$ sending each element $x$ of $X$ to $(\gamma, x)$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{H})$.
The lemma follows.
\end{proof}
\begin{lem}\label{lem-qn-res}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $\mathcal{S}$ be a quasi-normal subgroupoid of $\mathcal{G}$.
Then for any Borel subset $A$ of $X$ with positive measure, $(\mathcal{S})_A$ is quasi-normal in $(\mathcal{G})_A$.
\end{lem}
\begin{proof}
Let $\{ \phi_n\}_{n\in N}$ be a countable family of elements of ${\rm QN}_{\mathcal{G}}(\mathcal{S})$ such that for a.e.\ $g\in \mathcal{G}$, there exists $n\in N$ with $s(g)\in D_{\phi_n}$ and $\phi_n(s(g))g^{-1}\in \mathcal{S}$.
Choose a Borel map $\psi \colon \mathcal{S}A\to \mathcal{S}$ such that $\psi(x)=e_x$ for any $x\in A$, where $e_x$ is the unit element at $x$; and $s\circ \psi(x)=x$ and $r\circ \psi(x)\in A$ for any $x\in \mathcal{S}A\setminus A$.
For each $n\in N$, we set
\[D_n=\{\, x\in A\cap D_{\phi_n}\mid r\circ \phi_n(x)\in \mathcal{S}A\,\}\]
and define a Borel map $\psi_n\colon D_n\to (\mathcal{G})_A$ by $\psi_n(x)=\psi(r\circ \phi_n(x))\phi_n(x)$ for $x\in D_n$.
Taking a countable Borel partition of $D_n$ and restricting $\psi_n$ to each piece, we obtain a countable family $\{ \eta_m\}_{m\in M}$ of elements of ${\rm QN}_{(\mathcal{G})_A}((\mathcal{S})_A)$ such that for any $n\in N$ and a.e.\ $x\in D_n$, there exists $m\in M$ with $x\in D_{\eta_m}$ and $\eta_m(x)=\psi_n(x)$.
For a.e.\ $g\in (\mathcal{G})_A$, there exists $n\in N$ with $s(g)\in D_{\phi_n}$ and $\phi_n(s(g))g^{-1}\in \mathcal{S}$.
Putting $x=s(g)$, we have $r\circ \phi_n(x)\in \mathcal{S}A$.
We can thus find $m\in M$ with $x\in D_{\eta_m}$ and $\psi_n(x)=\eta_m(x)$.
The equality $\eta_m(x)g^{-1}=\psi(r\circ \phi_n(x))\phi_n(x)g^{-1}$ holds, and this element belongs to $(\mathcal{S})_A$.
It follows that $(\mathcal{S})_A$ is quasi-normal in $(\mathcal{G})_A$.
\end{proof}
\begin{lem}\label{lem-qn-finite}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $\mathcal{S}$ and $\mathcal{T}$ be subgroupoids of $\mathcal{G}$ such that $\mathcal{S}<\mathcal{T}$ and $[\mathcal{T}:\mathcal{S}]_x<\infty$ for a.e.\ $x\in X$.
Then we have the equality ${\rm QN}_{\mathcal{G}}(\mathcal{S})={\rm QN}_{\mathcal{G}}(\mathcal{T})$.
In particular, $\mathcal{S}$ is quasi-normal in $\mathcal{G}$ if and only if $\mathcal{T}$ is quasi-normal in $\mathcal{G}$.
\end{lem}
\begin{proof}
We denote by $r, s\colon \mathcal{G}\to X$ the range and source maps of $\mathcal{G}$, respectively.
Pick $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$.
For any $x\in R_{\phi}$, let $E$ be a set of representatives of all classes in $(s^{-1}(x)\cap (\mathcal{T})_{R_{\phi}})/(\mathcal{S})_{R_{\phi}}$.
We have
\[[(\mathcal{T})_{R_{\phi}}: (\mathcal{T})_{R_{\phi}}\cap \mathcal{T}^{\phi}]_x\leq [(\mathcal{T})_{R_{\phi}}: (\mathcal{S})_{R_{\phi}}\cap \mathcal{S}^{\phi}]_x= \sum_{g\in E}[(\mathcal{S})_{R_{\phi}}: (\mathcal{S})_{R_{\phi}}\cap \mathcal{S}^{\phi}]_{r(g)},\]
where the last equality holds by Lemma \ref{lem-index-formula} (iii).
For a.e.\ $x\in R_{\phi}$, the right hand side is finite because $E$ is finite and $\phi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
As in the last part in the proof of Lemma \ref{lem-qn} (ii), we can conclude that $[\mathcal{T}^{\phi}: (\mathcal{T})_{R_{\phi}}\cap \mathcal{T}^{\phi}]_x$ is also finite for a.e.\ $x\in R_{\phi}$.
It follows that $\phi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{T})$.
Pick $\psi \in {\rm QN}_{\mathcal{G}}(\mathcal{T})$.
For any $x\in R_{\psi}$, let $F$ be a set of representatives of all classes in $(s^{-1}(x)\cap (\mathcal{T})_{R_{\psi}})/((\mathcal{T})_{R_{\psi}}\cap \mathcal{T}^{\psi})$.
We have
\begin{align*}
& [(\mathcal{S})_{R_{\psi}}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_x\leq [(\mathcal{T})_{R_{\psi}}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_x\\
= \, & \sum_{g\in F}[(\mathcal{T})_{R_{\psi}}\cap \mathcal{T}^{\psi}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_{r(g)}.
\end{align*}
For a.e.\ $x\in R_{\psi}$, the set $F$ is finite because $\psi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{T})$.
For any $g\in F$, putting $y=r(g)$, we have
\begin{align*}
& [(\mathcal{T})_{R_{\psi}}\cap \mathcal{T}^{\psi}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_y\\
\leq \, & [(\mathcal{T})_{R_{\psi}}\cap \mathcal{T}^{\psi}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{T}^{\psi}]_y[(\mathcal{T})_{R_{\psi}}\cap \mathcal{T}^{\psi}: \mathcal{S}^{\psi}\cap (\mathcal{T})_{R_{\psi}}]_y\\
\leq \, & [(\mathcal{T})_{R_{\psi}}: (\mathcal{S})_{R_{\psi}}]_y[\mathcal{T}^{\psi}: \mathcal{S}^{\psi}]_y,
\end{align*}
where the first and second inequalities hold by Lemma \ref{lem-index-formula} (i) and (ii), respectively.
The right hand side is finite for a.e.\ $x\in R_{\psi}$.
We see that $[(\mathcal{S})_{R_{\psi}}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_x$ is finite for a.e.\ $x\in R_{\psi}$.
Similarly, we can conclude that $[\mathcal{S}^{\psi}: (\mathcal{S})_{R_{\psi}}\cap \mathcal{S}^{\psi}]_x$ is finite for a.e.\ $x\in R_{\psi}$.
It follows that $\psi$ belongs to ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
\end{proof}
\subsection{Quotient}
Let $(X, \mu)$ be a standard finite measure space.
Let $\mathcal{G}$ be a discrete measured groupoid on $(X, \mu)$.
Given a normal subgroupoid $\mathcal{S}$ of $\mathcal{G}$, we can construct a discrete measured groupoid $\mathcal{Q}$ on a standard finite measure space $(Z, \xi)$ and a Borel homomorphism $\theta \colon \mathcal{G}\to \mathcal{Q}$ satisfying the following three conditions:
\begin{enumerate}
\item[(a)] The equality $\ker \theta =\mathcal{S}$ holds.
\item[(b)] For a.e.\ $h\in \mathcal{Q}$ and $x\in X$ such that $\theta(x)$ is equal to the source of $h$, there exists $g\in \mathcal{G}$ with $s(g)=x$ and $\theta(g)=h$, where the map from $X$ into $Z$ induced by $\theta$ is denoted by the same symbol $\theta$.
\item[(c)] If $\mathcal{Q}'$ is a discrete measured groupoid on a standard finite measure space $(Z', \xi')$ and if $\theta'\colon \mathcal{G}\to \mathcal{Q}'$ is a Borel homomorphism with $\mathcal{S}<\ker \theta'$, then there exists a Borel homomorphism $\tau \colon \mathcal{Q}\to \mathcal{Q}'$ with $\tau \circ \theta =\theta'$.
\end{enumerate}
The groupoid $\mathcal{Q}$ is called the {\it quotient} of $\mathcal{G}$ by $\mathcal{S}$ and denoted by $\mathcal{G}/\mathcal{S}$.
We refer to the proof of \cite[Theorem 2.2]{fsz} for the construction of $\mathcal{Q}$ and $\theta$ (see also \cite[Section 3]{sauer-thom}, where the quotient by a strongly normal subgroupoid is discussed).
Although in \cite{fsz}, the quotient is constructed in the case where $\mathcal{G}$ is principal, it is also valid in the general case.
In the construction, the map from $(X, \mu)$ into $(Z, \xi)$ induced by $\theta$ is defined as the ergodic decomposition for $\mathcal{S}$.
In particular, if $\mathcal{S}$ is ergodic, then $\mathcal{Q}$ is a discrete group.
The following lemma is deduced from the construction of the quotient.
\begin{lem}\label{lem-qu}
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $\mathcal{H}$ be a normal subgroupoid of $\mathcal{G}$.
Then the following assertions hold:
\begin{enumerate}
\item Let $A$ be a Borel subset of $X$ with $X=\mathcal{H}A$.
Then the inclusion of $(\mathcal{G})_A$ into $\mathcal{G}$ induces an isomorphism from $(\mathcal{G})_A/(\mathcal{H})_A$ onto $\mathcal{G}/\mathcal{H}$.
\item Suppose that we have a non-singular action of a discrete group $\Gamma$ on $(X, \mu)$ and a normal subgroup $\Lambda$ of $\Gamma$ such that $\mathcal{G}=\Gamma \ltimes (X, \mu)$ and $\mathcal{H}=\Lambda \ltimes (X, \mu)$.
If the action of $\Lambda$ on $(X, \mu)$ is ergodic, then the projection from $\mathcal{G}$ onto $\Gamma$ induces an isomorphism from $\mathcal{G}/\mathcal{H}$ onto $\Gamma/\Lambda$.
\end{enumerate}
\end{lem}
\section{A variant of Furman's theorem}\label{sec-var}
Theorem \ref{thm-furman} below is used to construct a representation of a group ME to a given group.
It is proved by Furman \cite{furman-mer} in the framework of higher rank lattices, and plays a significant role to deduce ME and OE rigidity results in \cite{bfs}, \cite{furman-mer}, \cite{furman-oer}, \cite{kida-oer} and \cite{ms}, etc.
We provide a variant of this theorem to get a representation of a group ME to a Baumslag-Solitar group, into $\mathbb{R}$.
We first review measure equivalence and introduce terminology.
\begin{defn}[\ci{0.5.E}{gromov-as-inv}]\label{defn-me}
Two discrete groups $\Gamma$ and $\Lambda$ are said to be {\it measure equivalent (ME)} if we have a standard Borel space $(\Sigma, m)$ with a $\sigma$-finite positive measure and a measure-preserving action of $\Gamma \times \Lambda$ on $(\Sigma, m)$ such that there exist Borel subsets $X, Y\subset \Sigma$ satisfying $m(X)<\infty$, $m(Y)<\infty$ and the equality
\[\Sigma =\bigsqcup_{\gamma \in \Gamma}(\gamma, e)Y=\bigsqcup_{\lambda \in \Lambda}(e, \lambda)X\]
up to $m$-null sets.
The space $(\Sigma, m)$ equipped with the action of $\Gamma \times \Lambda$ is then called a {\it $(\Gamma, \Lambda)$-coupling}.
\end{defn}
ME is an equivalence relation between discrete groups (see \cite[Section 2]{furman-mer}).
It is known that two discrete groups $\Gamma$ and $\Lambda$ are ME if and only if there exists an ergodic f.f.m.p.\ action of $\Gamma$ which is WOE to an ergodic f.f.m.p.\ action of $\Lambda$, as discussed in \cite[Section 3]{furman-oer}.
Let $\Gamma$ and $\Lambda$ be discrete groups, and let $G$ be a standard Borel group.
Given homomorphisms $\pi \colon \Gamma \to G$ and $\rho \colon \Lambda \to G$, we denote by $(G, \pi, \rho)$ the Borel space $G$ equipped with the action of $\Gamma \times \Lambda$ on it defined by
\[(\gamma, \lambda)g=\pi(\gamma)g\rho(\lambda)^{-1},\quad g\in G,\ \gamma \in \Gamma,\ \lambda \in \Lambda.\]
Let $\Sigma$ be a $(\Gamma, \Lambda)$-coupling, and let $\Phi \colon \Sigma \to S$ be a Borel map into a standard Borel space $S$ on which $\Gamma \times \Lambda$ acts.
We say that $\Phi$ is {\it almost $(\Gamma \times \Lambda)$-equivariant} if we have the equality
\[\Phi((\gamma, \lambda)x)=(\gamma, \lambda)\Phi(x),\quad \forall \gamma \in \Gamma,\ \forall \lambda \in \Lambda,\ \textrm{a.e.}\ x\in \Sigma.\]
For a $(\Gamma, \Lambda)$-coupling $\Sigma$, we denote by $\Sigma \times_{\Lambda}\Sigma$ the quotient space of $\Sigma \times \Sigma$ by the diagonal action of $\Lambda$, which is a $(\Gamma, \Gamma)$-coupling.
We refer to \cite[Theorem 2.5]{bfs} for the proof of the following:
\begin{thm}\label{thm-furman}
Let $\Gamma$ be a discrete group, $G$ a standard Borel group and $\pi \colon \Gamma \to G$ a homomorphism.
Let $\Lambda$ be a discrete group, and let $(\Sigma, m)$ be a $(\Gamma, \Lambda)$-coupling.
We set $\Omega =\Sigma \times_{\Lambda}\Sigma$.
Suppose that
\begin{enumerate}
\item[(a)] the Dirac measure on the neutral element of $G$ is the only probability measure on $G$ invariant under conjugation by any element of $\pi(\Gamma)$; and
\item[(b)] we have an almost $(\Gamma \times \Gamma)$-equivariant Borel map from $\Omega$ into $(G, \pi, \pi)$.
\end{enumerate}
Then there exist a homomorphism $\rho \colon \Lambda \rightarrow G$ and an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \rightarrow (G, \pi, \rho)$.
In addition, if $\ker \pi$ is finite and there is a Borel fundamental domain for the action of $\pi(\Gamma)$ on $G$ by left multiplication, then $\rho$ can be chosen so that $\ker \rho$ is finite.
\end{thm}
We say that a measure-preserving action of a discrete group $\Gamma$ on a probability space $(X, \mu)$ is {\it weakly mixing} if the diagonal action of $\Gamma$ on the product $(X\times X, \mu \times \mu)$ is ergodic.
This condition is known to imply that for any ergodic measure-preserving action of $\Gamma$ on a probability space $(Y, \nu)$, the diagonal action of $\Gamma$ on the product $(X\times Y, \mu \times \nu)$ is ergodic (see \cite[Proposition 2.2]{sch}).
We note that for any $(\Gamma, \Lambda)$-coupling $\Sigma$, the action $\Gamma \times \Gamma \c \Sigma \times_{\Lambda}\Sigma$ is ergodic if and only if the action $\Lambda \c \Sigma /\Gamma$ is weakly mixing.
\begin{thm}\label{thm-abel}
Let $\Gamma$ be a discrete group, $G$ an abelian standard Borel group and $\pi \colon \Gamma \to G$ a homomorphism.
Let $\Lambda$ be a discrete group, and let $(\Sigma, m)$ be a $(\Gamma, \Lambda)$-coupling.
We set $\Omega =\Sigma \times_{\Lambda}\Sigma$.
Suppose that
\begin{enumerate}
\item[(1)] the action $\Lambda \c \Sigma /\Gamma$ is weakly mixing; and
\item[(2)] we have an almost $(\Gamma \times \Gamma)$-equivariant Borel map from $\Omega$ into $(G, \pi, \pi)$.
\end{enumerate}
Then there exist a homomorphism $\rho \colon \Lambda \rightarrow G$ and an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to (G, \pi, \rho)$.
\end{thm}
\begin{proof}
For $(x, y)\in \Sigma \times \Sigma$, we denote by $[x, y]\in \Omega$ the equivalence class of $(x, y)$.
Let $\Psi \colon \Omega \to (G, \pi, \pi)$ be an almost $(\Gamma \times \Gamma)$-equivariant Borel map.
\begin{claim}\label{claim-h}
Define a Borel map $H\colon \Sigma^4\to G$ by
\[H(x, y, z, w)=\Psi([y, z])\Psi([x, z])^{-1}\Psi([x, w])\Psi([y, w])^{-1}\]
for $(x, y, z, w)\in \Sigma^4$.
Then $H$ is essentially constant.
\end{claim}
\begin{proof}
For any $\gamma \in \Gamma$, $\lambda \in \Lambda$ and $(x, y, z, w)\in \Sigma^4$, we have
\begin{align*}
H(x, y, z, w)&=H(\gamma x, y, z, w)=H(x, \gamma y, z, w)=H(x, y, \gamma z, w)=H(x, y, z, \gamma w)\\
&=H(\lambda x, \lambda y, \lambda z, \lambda w)
\end{align*}
because $G$ is abelian.
It follows that $H$ induces a Borel map $\bar{H}\colon (\Sigma /\Gamma)^4\to G$ which is invariant under the diagonal action of $\Lambda$ on $(\Sigma /\Gamma)^4$.
This action of $\Lambda$ is ergodic because the action $\Lambda \c \Sigma /\Gamma$ is weakly mixing.
The map $\bar{H}$ is therefore essentially constant, and so is $H$.
\end{proof}
Let $g_0\in G$ denote the essential value of the map $H$.
For $z\in \Sigma$, we define a Borel map $F_z\colon \Sigma^2\to G$ by
\[F_z(x, y)=\Psi([x, z])\Psi([y, z])^{-1}\]
for $(x, y)\in \Sigma^2$.
By Claim \ref{claim-h} and Fubini's theorem, for a.e.\ $x\in \Sigma$ and any $\lambda \in \Lambda$, we have
\[F_z(x, y)^{-1}F_w(x, y)=g_0=F_z(\lambda^{-1}x, y)^{-1}F_w(\lambda^{-1}x, y)\]
for a.e.\ $(y, z, w)\in \Sigma^3$.
For a.e.\ $x\in \Sigma$ and any $\lambda \in \Lambda$, the equality
\[F_z(\lambda^{-1}x, y)F_z(x, y)^{-1}=\Psi([\lambda^{-1}x, z])\Psi([x, z])^{-1}\]
for any $y\in \Sigma$ implies that the Borel map $\Sigma \ni z\mapsto \Psi([\lambda^{-1}x, z])\Psi([x, z])^{-1}\in G$ is essentially constant.
We define $\rho_x(\lambda)\in G$ to be the essential value of this map.
For a.e.\ $x\in \Sigma$ and any $\lambda_1, \lambda_2\in \Lambda$, choosing some $z\in \Sigma$, we have
\begin{align*}
\rho_x(\lambda_1\lambda_2^{-1})&=\Psi([\lambda_2\lambda_1^{-1}x, z])\Psi([x, z])^{-1}=\Psi([\lambda_1^{-1}x, \lambda_2^{-1}z])\Psi([x, z])^{-1}\\
&=\Psi([\lambda_1^{-1}x, \lambda_2^{-1}z])\Psi([x, \lambda_2^{-1}z])^{-1}\Psi([x, \lambda_2^{-1}z])\Psi([\lambda_2^{-1}x, \lambda_2^{-1}z])^{-1}\\
&=\rho_x(\lambda_1)\rho_x(\lambda_2)^{-1}.
\end{align*}
For a.e.\ $x\in \Sigma$, the map $\rho_x\colon \Lambda \to G$ is therefore a homomorphism.
There exists an element $x_0$ of $\Sigma$ such that $\rho_{x_0}\colon \Lambda \to G$ is a homomorphism and we have the equality $\rho_{x_0}(\lambda)=\Psi([\lambda^{-1}x_0, x])\Psi([x_0, x])^{-1}$ for any $\lambda \in \Lambda$ and a.e.\ $x\in \Sigma$.
We define a Borel map $\Phi \colon \Sigma \to G$ by $\Phi(x)=\Psi([x_0, x])^{-1}$ for $x\in \Sigma$.
The map $\Phi \colon \Sigma \to (G, \pi, \rho)$ is then almost $(\Gamma \times \Lambda)$-equivariant.
\end{proof}
\section{Elliptic subgroupoids}\label{sec-ell}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$.
Let $T$ be the Bass-Serre tree associated with $\Gamma$.
Suppose that we have a measure-preserving action of $\Gamma$ on a standard finite measure space $(X, \mu)$.
We set $\mathcal{G}=\Gamma \ltimes X$ and define a homomorphism $\rho \colon \mathcal{G}\to \Gamma$ by $\rho(\gamma, x)=\gamma$ for $(\gamma, x)\in \mathcal{G}$.
Let $A$ be a Borel subset of $X$ with positive measure.
We say that a subgroupoid $\mathcal{S}$ of $(\mathcal{G})_A$ is {\it elliptic} if there exists an $(\mathcal{S}, \rho)$-invariant Borel map from $A$ into $V(T)$.
For any $v\in V(T)$, the subgroupoid $\Gamma_v\ltimes X$ of $\mathcal{G}$ is of infinite type, amenable and quasi-normal in $\mathcal{G}$ because $\Gamma_v$ is quasi-normal in $\Gamma$.
Conversely, Theorem \ref{thm-ell} below says that these algebraic properties imply ellipticity.
Let us say that a discrete measured groupoid $\mathcal{H}$ on a standard finite measure space $(Y, \nu)$ is {\it nowhere amenable} if for any Borel subset $B$ of $Y$ with positive measure, $(\mathcal{H})_B$ is not amenable.
\begin{thm}\label{thm-ell}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$.
Suppose that we have a measure-preserving action of $\Gamma$ on a standard finite measure space $(X, \mu)$, and set $\mathcal{G}=\Gamma \ltimes X$.
Let $A$ be a Borel subset of $X$ with positive measure, and let $\mathcal{S}$ and $\mathcal{T}$ be subgroupoids of $(\mathcal{G})_A$ such that we have $\mathcal{S}<\mathcal{T}$; $\mathcal{S}$ is amenable and is quasi-normal in $\mathcal{T}$; and $\mathcal{T}$ is nowhere amenable.
Then $\mathcal{S}$ is elliptic.
\end{thm}
Before proving this theorem, we prepare the following:
\begin{notation}\label{not-tree}
For a locally compact Polish space $K$, we denote by $M(K)$ the space of probability measures on $K$ equipped with the weak* topology.
Let $V_f(T)$ denote the set of non-empty finite subsets of $V(T)$.
Let $S(T)$ denote the set of simplices of $T$.
We have the ${\rm Aut}(T)$-equivariant map $C\colon V_f(T)\to S(T)$ associating to each element of $V_f(T)$ its barycenter (see \cite[Section 4.2]{kida-exama} for a precise definition).
We also have the ${\rm Aut}(T)$-equivariant Borel map $M\colon M(V(T))\to V_f(T)$ associating to each $\nu \in M(V(T))$ the set of all elements of $V(T)$ attaining the maximal value of the function $\nu$ on $V(T)$.
We set
\[\delta T=\{\, (x, y, z)\in (\partial T)^3\mid x\neq y\neq z\neq x\,\},\]
where $\Gamma$ acts by the formula $\gamma (x, y, z)=(\gamma x, \gamma y, \gamma z)$ for $\gamma \in \Gamma$ and $(x, y, z)\in \delta T$.
We define a $\Gamma$-equivariant Borel map $G\colon \delta T\rightarrow V(T)$ so that for each $(x, y, z)\in \delta T$, $G(x, y, z)$ is the intersection of the three geodesics in $T$ joining two of $x$, $y$ and $z$.
We define $\partial_2 T$ as the quotient of $\partial T\times \partial T$ by the action of the symmetric group of two letters that exchanges the coordinates.
\end{notation}
\begin{proof}[Proof of Theorem \ref{thm-ell}]
This proof is similar to the proof of \cite[Lemmas 4.2, 4.4 and 4.5]{kida-exama} except for using quasi-normality in place of normality.
The argument in \cite{kida-exama} partially depends on the proof of \cite[Lemma 3.2]{adams} (see also \cite[Chapter 2]{hjorth-kechris}).
Let $\rho \colon \mathcal{G}\to \Gamma$ be the homomorphism defined by $\rho(\gamma, x)=\gamma$ for $(\gamma, x)\in \mathcal{G}$.
For a Borel subset $B$ of $A$ with positive measure, we define $I_B$ as the set of all $(\mathcal{S}, \rho)$-invariant Borel maps from $B$ into $V(T)$.
It is enough to deduce a contradiction under the assumption that there exists a Borel subset $B$ of $A$ with positive measure such that for any Borel subset $B_1$ of $B$ with positive measure, $I_{B_1}$ is empty.
Since $\mathcal{S}$ is amenable, there exists an $(\mathcal{S}, \rho)$-invariant Borel map from $A$ into $M(\partial T)$.
\begin{lem}\label{lem-two}
Let $B_1$ be a Borel subset of $B$ with positive measure, and let $\varphi \colon B_1\to M(\partial T)$ be an $(\mathcal{S}, \rho)$-invariant Borel map.
Then for a.e.\ $x\in B_1$, the measure $\varphi(x)$ is supported on at most two points of $\partial T$.
\end{lem}
\begin{proof}
If the lemma were not true, then there would exist a Borel subset $B_2$ of $B_1$ with positive measure such that for a.e.\ $x\in B_2$, the restriction of the measure $\varphi(x)^3$ on $(\partial T)^3$ to $\delta T$ is non-zero.
Composing the map assigning to each $x\in B_2$ the normalization of the restriction of $\varphi(x)^3$ to $\delta T$, we obtain an $(\mathcal{S}, \rho)$-invariant Borel map from $B_2$ into $M(\delta T)$.
Composing the maps $G$, $M$ and $C$, we obtain an $(\mathcal{S}, \rho)$-invariant Borel map from $B_2$ into $S(T)$.
We also obtain an $(\mathcal{S}, \rho)$-invariant Borel map from $B_2$ into $V(T)$ because $\Gamma$ acts on $T$ without inversions.
This contradicts our assumption that $I_{B_2}$ is empty.
\end{proof}
Lemma \ref{lem-two} implies that any $(\mathcal{S}, \rho)$-invariant Borel map $\varphi \colon B\to M(\partial T)$ induces an $(\mathcal{S}, \rho)$-invariant Borel map from $B$ into $\partial_2T$.
We define $J$ as the set of all $(\mathcal{S}, \rho)$-invariant Borel maps from $B$ into $\partial_2T$, which is non-empty.
For each $\varphi \in J$, we set
\[S_{\varphi}=\{\, x\in B\mid |{\rm supp}(\varphi(x))|=2\,\},\]
where ${\rm supp}(\nu)$ denotes the support of a measure $\nu$.
There exists an element $\varphi_0$ of $J$ with $\mu(S_{\varphi_0})=\sup_{\varphi \in J}\mu(S_{\varphi})$.
\begin{lem}\label{lem-fi-inv}
In the above notation, let $B_1$ be a Borel subset of $B$ with positive measure.
Let $\mathcal{S}_0$ be a subgroupoid of $(\mathcal{S})_{B_1}$ with $[(\mathcal{S})_{B_1}: \mathcal{S}_0]_x<\infty$ for a.e.\ $x\in B_1$.
Then for any $(\mathcal{S}_0, \rho)$-invariant Borel map $\varphi \colon B_1\to \partial_2 T$, we have $\varphi(x)\subset \varphi_0(x)$ for a.e.\ $x\in B_1$.
\end{lem}
\begin{proof}
We naturally identify $\partial_2 T$ with a subset of $\mathcal{F}(\partial T)$, the space of non-empty finite subsets of $\partial T$.
Let $\varphi \colon B_1\to \partial_2 T$ be an $(\mathcal{S}_0, \rho)$-invariant Borel map.
By Lemma \ref{lem-finite-index}, there exists an $(\mathcal{S}, \rho)$-invariant Borel map $\Phi \colon B_1\to \mathcal{F}(\partial T)$ such that $\varphi(x)\subset \Phi(x)$ for a.e.\ $x\in B_1$.
Since each element of $\mathcal{F}(\partial T)$ naturally associates a probability measure on $\partial T$, applying Lemma \ref{lem-two}, we have $\Phi(x)\in \partial_2 T$ for a.e.\ $x\in B_1$.
By the maximality of $\mu(S_{\varphi_0})$, we have $\Phi(x)\subset \varphi_0(x)$ for a.e.\ $x\in B_1$.
It thus turns out that $\varphi(x)\subset \varphi_0(x)$ for a.e.\ $x\in B_1$.
The lemma is proved.
\end{proof}
We claim that the map $\varphi_0$ is $(\mathcal{T}, \rho)$-invariant.
Pick $\psi \in {\rm QN}_{(\mathcal{T})_B}((\mathcal{S})_B)$.
It suffices to show that for a.e.\ $x\in D_{\psi}$, we have the equality $\rho(\psi(x))\varphi_0(x)=\varphi_0(r\circ \psi(x))$.
We define a Borel map $\chi \colon D_{\psi}\to \partial_2 T$ by $\chi(x)=\rho(\psi(x))^{-1}\varphi_0(r\circ \psi(x))$ for $x\in D_{\psi}$.
For a.e.\ $g\in (\mathcal{S})_{D_{\psi}}\cap U_{\psi}^{-1}((\mathcal{S})_{R_{\psi}})$ with $x=s(g)$ and $y=r(g)$, we have
\[\rho(g)\chi(x)=\rho(\psi(y))^{-1}\rho(\psi(y)g\psi(x)^{-1})\varphi_0(r\circ \psi(x))=\rho(\psi(y))^{-1}\varphi_0(r\circ \psi(y))=\chi(y),\]
where the second equality holds because $U_{\psi}(g)=\psi(y)g\psi(x)^{-1}$ belongs to $\mathcal{S}$.
Since $\psi$ is in ${\rm QN}_{(\mathcal{T})_B}((\mathcal{S})_B)$, we have $[(\mathcal{S})_{D_{\psi}}: (\mathcal{S})_{D_{\psi}}\cap U_{\psi}^{-1}((\mathcal{S})_{R_{\psi}})]_x<\infty$ for a.e.\ $x\in D_{\psi}$.
By Lemma \ref{lem-fi-inv}, the inclusion $\chi(x)\subset \varphi_0(x)$ holds for a.e.\ $x\in D_{\psi}$.
Let $\zeta \in {\rm QN}_{(\mathcal{T})_B}((\mathcal{S})_B)$ be the inverse of $\psi$.
We next define a Borel map $\omega \colon D_{\zeta}\to \partial_2 T$ by $\omega(y)=\rho(\zeta(y))^{-1}\varphi_0(r\circ \zeta(y))$ for $y\in D_{\zeta}$.
The argument in the last paragraph implies the inclusion $\omega(y)\subset \varphi_0(y)$ for a.e.\ $y\in D_{\zeta}$.
For a.e.\ $x\in D_{\psi}$, we have
\begin{align*}
\chi(x)&=\rho(\psi(x))^{-1}\varphi_0(r\circ \psi(x))\subset \varphi_0(x)=\rho(\psi(x))^{-1}\omega(r\circ \psi(x))\\
&\subset \rho(\psi(x))^{-1}\varphi_0(r\circ \psi(x))=\chi(x)
\end{align*}
and thus $\chi(x)=\varphi_0(x)$.
The claim is proved.
The action of $\Gamma$ on $\partial_2T$ is amenable in a measure-theoretic sense by \cite[Corollary 3.4]{kida-exama}.
The existence of the $(\mathcal{T}, \rho)$-invariant Borel map $\varphi_0\colon B\to \partial_2T$ therefore implies that $(\mathcal{T})_B$ is amenable by \cite[Proposition 2.5]{kida-exama}.
This is a contradiction because $\mathcal{T}$ is nowhere amenable.
\end{proof}
In the following theorem, we obtain a result similar to Theorem \ref{thm-ell} for a discrete group having an infinite amenable normal subgroup with the quotient hyperbolic.
\begin{thm}\label{thm-ell-hyp}
Let $\Delta$ be a discrete group, and let $N$ be an infinite, amenable and normal subgroup of $\Delta$ such that $\Delta /N$ is non-elementarily hyperbolic.
Suppose that we have a measure-preserving action of $\Delta$ on a standard finite measure space $(X, \mu)$.
We set $\mathcal{G}=\Delta \ltimes X$.
Let $A$ be a Borel subset of $X$ with positive measure, and let $\mathcal{S}$ and $\mathcal{T}$ be subgroupoids of $(\mathcal{G})_A$ such that we have $\mathcal{S}<\mathcal{T}$; $\mathcal{S}$ is amenable and is quasi-normal in $\mathcal{T}$; and $\mathcal{T}$ is nowhere amenable.
Then there exist a countable Borel partition $A=\bigsqcup_n A_n$ and a subgroup $L_n$ of $\Delta$ such that for each $n$, we have $N<L_n$, $[L_n: N]<\infty$ and $(\mathcal{S})_{A_n}<(L_n\ltimes X)_{A_n}$.
\end{thm}
\begin{proof}
The proof is essentially the same as that of Theorem \ref{thm-ell}.
We put $Q=\Delta /N$ and denote by $\partial Q$ the boundary of $Q$ as a hyperbolic metric space.
Let $Q$ act on itself by left multiplication.
This action of $Q$ extends to a continuous action on the compactification $Q\cup \partial Q$ of $Q$.
As before, for a locally compact Polish space $K$, we denote by $M(K)$ the space of probability measures on $K$ equipped with the weak* topology.
Let $Q_f$ denote the set of non-empty finite subsets of $Q$.
We have the $Q$-equivariant Borel map $M\colon M(Q)\to Q_f$ associating to each $\nu \in M(Q)$ the set of all elements of $Q$ attaining the maximal value of the function $\nu$ on $Q$.
We set
\[\delta Q =\{\, (x, y, z)\in (\partial Q)^3\mid x\neq y\neq z\neq x\,\},\]
where $Q$ acts by the formula $\gamma (x, y, z)=(\gamma x, \gamma y, \gamma z)$ for $\gamma \in Q$ and $(x, y, z)\in \delta Q$.
The set $\delta Q$ is non-empty because $Q$ is non-elementary.
In \cite[Definition 6.6]{adams-hyp}, Adams constructs a $Q$-equivariant Borel map $MS\colon \delta Q \to Q_f$, using hyperbolicity of $Q$.
We define $\partial_2Q$ as the quotient of $\partial Q \times \partial Q$ by the action of the symmetric group of two letters that exchanges the coordinates.
Let $\rho \colon \mathcal{G}\to Q$ be the composition of the projection from $\mathcal{G}$ onto $\Delta$ with the quotient map from $\Delta$ onto $Q$.
For a Borel subset $B$ of $A$ with positive measure, we define $I_B$ as the set of all $(\mathcal{S}, \rho)$-invariant Borel maps from $B$ into $Q_f$.
Assume that there is a Borel subset $B$ of $A$ with positive measure such that for any Borel subset $B_1$ of $B$ with positive measure, $I_{B_1}$ is empty.
Along the proof of Theorem \ref{thm-ell}, we can deduce a contradiction, using the fact that the action of $Q$ on $\partial_2Q$ is amenable in a measure-theoretic sense, proved in \cite[Theorem 5.1]{adams-hyp}.
It follows that $I_A$ is non-empty, that is, there exists an $(\mathcal{S}, \rho)$-invariant Borel map $\psi \colon A\to Q_f$.
Take a countable Borel partition $A=\bigsqcup_n A_n$ such that for each $n$, the map $\psi$ is constant on $A_n$.
Since the stabilizer of each element of $Q_f$ in $Q$ is finite, for each $n$, there exists a subgroup $L_n$ of $\Delta$ such that $N<L_n$, $[L_n: N]<\infty$ and $(\mathcal{S})_{A_n}<(L_n\ltimes X)_{A_n}$.
\end{proof}
\section{The modular cocycles}\label{sec-mod}
Given a finite-measure-preserving, discrete measured groupoid $\mathcal{G}$ and its quasi-normal subgroupoid $\mathcal{S}$, we introduce the modular cocycle of Radon-Nikodym type and the local-index cocycle in Sections \ref{subsec-rn} and \ref{subsec-lic}, respectively.
They are defined by measuring a difference between $\mathcal{S}$ and its conjugate by an element of ${\rm QN}_{\mathcal{G}}(\mathcal{S})$.
In Section \ref{subsec-comp}, these two cocycles are computed when $\mathcal{G}$ and $\mathcal{S}$ are associated with an action of a Baumslag-Solitar group and its restriction to an elliptic subgroup.
\subsection{The modular cocycle of Radon-Nikodym type}\label{subsec-rn}
In this subsection, unless otherwise stated, let $\mathcal{G}$ be a measure-preserving, discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $\mathcal{S}$ be a quasi-normal subgroupoid of $\mathcal{G}$.
Fix $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$ with $\mu(D_{\phi})>0$.
We first define $\mathfrak{D}(\phi, x)\in \mathbb{R}_+^{\times}$ for $x\in D_{\phi}$.
We will define $\mathfrak{D} \colon \mathcal{G}\to \mathbb{R}_+^{\times}$, called the modular cocycle of Radon-Nikodym type for $\mathcal{G}$ and $\mathcal{S}$, so that $\mathfrak{D}(\phi(x))=\mathfrak{D}(\phi, x)$ for a.e.\ $x\in D_{\phi}$.
We put $D=D_{\phi}$, $R=R_{\phi}$ and $U=U_{\phi}$.
We also put
\[\mathcal{S}_-=(\mathcal{S})_D\cap U^{-1}((\mathcal{S})_R),\quad \mathcal{S}_+=(\mathcal{S})_R\cap U((\mathcal{S})_D).\]
The restriction of $U$ is an isomorphism from $\mathcal{S}_-$ onto $\mathcal{S}_+$.
Unless there is a confusion, for each $x\in D$, we write $U(x)=r\circ \phi(x)\in R$.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ be the ergodic decomposition for $\mathcal{S}$.
Let $\mu =\int_Z \mu_z d\xi(z)$ be the disintegration with respect to $\pi$.
For $z\in Z$, we put $X_z=\pi^{-1}(z)$.
Let
\[\pi_-\colon (D, \mu|_D)\to (Z_-, \xi_-)\quad \textrm{and}\quad \pi_+\colon (R, \mu|_R)\to (Z_+, \xi_+)\]
be the ergodic decompositions for $\mathcal{S}_-$ and $\mathcal{S}_+$, respectively, with $(\pi_-)_*(\mu|_D)=\xi_-$ and $(\pi_+)_*(\mu|_R)=\xi_+$.
For $z\in Z_-$, we put $D_z=\pi_-^{-1}(z)$, and for $w\in Z_+$, we put $R_w=\pi_+^{-1}(w)$.
Let $Y_-$ and $Y_+$ be the Borel subsets of $Z$ with $\mathcal{S}D=\pi^{-1}(Y_-)$ and $\mathcal{S}R=\pi^{-1}(Y_+)$.
We define $\pi_D\colon D\to Y_-$ and $\pi_R\colon R\to Y_+$ as the restrictions of $\pi$.
We set $\eta_-=(\pi_D)_*(\mu|_D)$ and $\eta_+=(\pi_R)_*(\mu|_R)$.
Let $\sigma_-\colon Z_-\to Y_-$ and $\sigma_+\colon Z_+\to Y_+$ be the canonical Borel maps such that the following diagrams commute:
\[\xymatrix{
& \ar[dl]_{\pi_-} (D, \mu|_D) \ar[dr]^{\pi_D} & \\
(Z_-, \xi_-) \ar[rr]^{\sigma_-} & & (Y_-, \eta_-) \\
}
\quad
\xymatrix{
& \ar[dl]_{\pi_+} (R, \mu|_R) \ar[dr]^{\pi_R} & \\
(Z_+, \xi_+) \ar[rr]^{\sigma_+} & & (Y_+, \eta_+) \\
}
\]
For $y\in Y_-$, we put $(Z_-)_y=\sigma_-^{-1}(y)$, and for $y\in Y_+$, we put $(Z_+)_y=\sigma_+^{-1}(y)$.
The map $\pi_D$ is the ergodic decomposition for $(\mathcal{S})_D$.
We have the disintegration
\[\mu|_D=\int_{Y_-}\mu_y(D\cap X_y)^{-1}\mu_y|_D\, d\eta_-(y)\]
with respect to $\pi_D$ because the equality $\eta_-=\chi\xi|_{Y_-}$ holds, where $\chi$ is the Borel function on $Y_-$ defined by $\chi(y)=\mu_y(D\cap X_y)$ for $y\in Y_-$.
Applying Lemma \ref{lem-erg-dec} to $(\mathcal{S})_D$ and $\mathcal{S}_-$, we may assume that for a.e.\ $y\in Y_-$, the set $(Z_-)_y$ is finite.
Applying Lemma \ref{lem-rest} to $\pi_D$, $\pi_-$ and $\sigma_-$, we see that for a.e.\ $z\in Z_-$, the restriction $\mu_{\sigma_-(z)}|_{D_z}$ is a constant multiple of the ergodic measure for $(\mathcal{S}_-)_z$.
In the same manner, we may assume that for a.e.\ $y\in Y_+$, the set $(Z_+)_y$ is finite, and we see that for a.e.\ $z\in Z_+$, the restriction $\mu_{\sigma_+(z)}|_{R_z}$ is a constant multiple of the ergodic measure for $(\mathcal{S}_+)_z$.
Since $U$ is an isomorphism from $\mathcal{S}_-$ onto $\mathcal{S}_+$, for a.e.\ $x\in D$, the two measures on the set $U(D_{\pi_-(x)})=R_{\pi_+(U(x))}$,
\[U_*(\mu_{\pi(x)}|_{D_{\pi_-(x)}})\quad \textrm{and}\quad \mu_{\pi(U(x))}|_{R_{\pi_+(U(x))}},\]
are a constant multiple of each other.
We define a number $\mathfrak{D}(\phi, x)=\mathfrak{D}(\mathcal{G}, \mathcal{S}, \phi, x)\in \mathbb{R}_+^{\times}$ by the equality
\[U_*(\mu_{\pi(x)}|_{D_{\pi_-(x)}})=\mathfrak{D}(\phi, x)\mu_{\pi(U(x))}|_{R_{\pi_+(U(x))}}.\]
\begin{lem}\label{lem-d-res}
If $A$ is a Borel subset of $D$ with $\mu(A)>0$, then we have the equality $\mathfrak{D}(\phi|_A, x)=\mathfrak{D}(\phi, x)$ for a.e.\ $x\in A$.
\end{lem}
\begin{proof}
Let $W_-$ be the Borel subset of $Z_-$ with $\pi_-^{-1}(W_-)=\mathcal{S}_- A$.
The map $\rho_-\colon A\to W_-$ defined as the restriction of $\pi_-$ is the ergodic decomposition for $(\mathcal{S}_-)_A$.
Putting $B=U(A)$, we define $W_+$ and $\rho_+\colon B\to W_+$ similarly.
For a.e.\ $x\in A$, restricting the equality in the definition of $\mathfrak{D}(\phi, x)$ to $B$, we obtain the equality
\[U_*(\mu_{\pi(x)}|_{D_{\pi_-(x)}\cap A})=\mathfrak{D}(\phi, x)\mu_{\pi(U(x))}|_{R_{\pi_+(U(x))}\cap B}.\]
Putting $A_z=\rho_-^{-1}(z)$ for $z\in W_-$ and $B_w=\rho_+^{-1}(w)$ for $w\in W_+$, we have $D_{\pi_-(x)}\cap A=A_{\rho_-(x)}$ and $R_{\pi_+(U(x))}\cap B=B_{\rho_+(U(x))}$ for a.e.\ $x\in A$.
The lemma follows.
\end{proof}
We define a Borel function $\mathfrak{D} =\mathfrak{D}(\mathcal{G}, \mathcal{S})\colon \mathcal{G}\to \mathbb{R}_+^{\times}$ as follows.
Pick a countable family $\Phi$ of elements of ${\rm QN}_{\mathcal{G}}(\mathcal{S})$ such that $\mu(D_{\phi})>0$ for any $\phi \in \Phi$ and the equality $\mathcal{G}=\bigsqcup_{\phi \in \Phi}\phi(D_{\phi})$ holds.
For $g\in \mathcal{G}$, choose $\phi \in \Phi$ with $g\in \phi(D_{\phi})$, and set $\mathfrak{D}(g)=\mathfrak{D}(\phi, s(g))$.
This definition of $\mathfrak{D}$ does not depend on the choice of $\Phi$ by Lemma \ref{lem-d-res}.
We call $\mathfrak{D}$ the {\it modular cocycle of Radon-Nikodym type} for $\mathcal{G}$ and $\mathcal{S}$.
\begin{lem}
The function $\mathfrak{D} \colon \mathcal{G}\to \mathbb{R}_+^{\times}$ defined above is a Borel cocycle, that is, it preserves products.
\end{lem}
\begin{proof}
Pick $\phi, \psi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$ and put $\eta =\psi \bullet \phi$, $D=D_{\eta}$, $R=R_{\eta}$ and $U=U_{\eta}$.
We assume $\mu(D)>0$ and the equality $D_{\psi}=R_{\phi}=U_{\phi}(D)$.
Put $L=U_{\phi}(D)$ and
\[\mathcal{S}_-=(\mathcal{S})_D\cap U^{-1}((\mathcal{S})_R),\quad \mathcal{S}_-'=(\mathcal{S})_D\cap U_{\phi}^{-1}((\mathcal{S})_L),\quad \mathcal{S}_-''=(\mathcal{S})_L\cap U_{\psi}^{-1}((\mathcal{S})_R).\]
For a.e.\ $x\in D$, since $[(\mathcal{S})_D: \mathcal{S}_-]_x$ is finite, the set $D\cap X_{\pi(x)}$ is decomposed into finitely many ergodic components for $\mathcal{S}_-$ by Lemma \ref{lem-erg-dec}.
Similarly, for a.e.\ $x\in D$, since $[(\mathcal{S})_D: \mathcal{S}_-']_x$ is finite, the set $D\cap X_{\pi(x)}$ is decomposed into finitely many ergodic components for $\mathcal{S}_-'$.
Let $Z_1$ be the Borel subset of $Z$ with $\mathcal{S}D=\pi^{-1}(Z_1)$.
For a.e.\ $z\in Z_1$, pick ergodic components $D_-$ and $D_-'$ for $\mathcal{S}_-$ and $\mathcal{S}_-'$, respectively, with $D_-, D_-'\subset X_z$ and $\mu_z(D_-\cap D_-')>0$.
For $\mu_z$-a.e.\ $x\in D_-'$, the equality
\[(U_{\phi})_*(\mu_{\pi(x)}|_{D_-'})=\mathfrak{D}(\phi, x)\mu_{\pi(U_{\phi}(x))}|_{U_{\phi}(D_-')}\]
holds by the definition of $\mathfrak{D}(\phi, x)$.
For a.e.\ $y\in L$, since $[(\mathcal{S})_L: \mathcal{S}_-'']_y$ is finite, the set $L\cap X_{\pi(y)}$ is decomposed into finitely many ergodic components for $\mathcal{S}_-''$.
Let $D_-''$ be an ergodic component for $\mathcal{S}_-''$ with $\mu_{z'}(U_{\phi}(D_-\cap D_-')\cap D_-'')>0$, where $z'$ is the point of $Z$ with $U_{\phi}(D_-')\subset X_{z'}$.
Such $z'$ exists because we have $U_{\phi}(\mathcal{S}_-')\subset (\mathcal{S})_L$.
For $\mu_{z'}$-a.e.\ $y\in D_-''$, the equality
\[(U_{\psi})_*(\mu_{\pi(y)}|_{D_-''})=\mathfrak{D}(\psi, y)\mu_{\pi(U_{\psi}(y))}|_{U_{\psi}(D_-'')}\]
holds by the definition of $\mathfrak{D}(\psi, y)$.
Combining the above two equalities, we obtain $\mathfrak{D}(\eta, x)=\mathfrak{D}(\psi, U_{\phi}(x))\mathfrak{D}(\phi, x)$ for $\mu_z$-a.e.\ $x\in U_{\phi}^{-1}(U_{\phi}(D_-\cap D_-')\cap D_-'')$.
\end{proof}
The following two lemmas will be used to compute the cocycle $\mathfrak{D}$.
\begin{lem}\label{lem-d-finite}
Let $\mathcal{S}$ and $\mathcal{T}$ be subgroupoids of $\mathcal{G}$ such that $\mathcal{S}<\mathcal{T}$; $[\mathcal{T}:\mathcal{S}]_x<\infty$ for a.e.\ $x\in X$; and $\mathcal{T}$ is quasi-normal in $\mathcal{G}$.
We set $\mathfrak{D}_{\mathcal{S}}=\mathfrak{D}(\mathcal{G}, \mathcal{S})$ and $\mathfrak{D}_{\mathcal{T}}=\mathfrak{D}(\mathcal{G}, \mathcal{T})$.
Then there exists a Borel map $\psi \colon X\to \mathbb{R}_+^{\times}$ with the equality $\mathfrak{D}_{\mathcal{S}}(g)=\psi(r(g))\mathfrak{D}_{\mathcal{T}}(g)\psi(s(g))^{-1}$ for a.e.\ $g\in \mathcal{G}$.
\end{lem}
\begin{proof}
Note that by Lemma \ref{lem-qn-finite}, $\mathcal{S}$ is quasi-normal in $\mathcal{G}$.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ and $\theta \colon (X, \mu)\to (W, \omega)$ be the ergodic decompositions for $\mathcal{S}$ and $\mathcal{T}$, respectively.
Let $\mu =\int_Z \mu_z d\xi(z)$ and $\mu =\int_W\nu_w d\omega(w)$ be the disintegrations with respect to $\pi$ and $\theta$, respectively.
We have the canonical Borel map $\tau \colon (Z, \xi)\to (W, \omega)$ with $\theta =\tau \circ \pi$.
Pick an element $\phi$ of the set ${\rm QN}_{\mathcal{G}}(\mathcal{T})={\rm QN}_{\mathcal{G}}(\mathcal{S})$ with $\mu(D_{\phi})>0$.
We put $D=D_{\phi}$, $R=R_{\phi}$ and $U=U_{\phi}$, and define $\mathcal{S}_-$, $\mathcal{S}_+$, $\pi_-$ and $\pi_+$ as in the definition of $\mathfrak{D}(\mathcal{G}, \mathcal{S})$.
Similarly, we set
\[\mathcal{T}_-=(\mathcal{T})_D\cap U^{-1}((\mathcal{T})_R),\quad \mathcal{T}_+=(\mathcal{T})_R\cap U((\mathcal{T})_D).\]
Let $\theta_-\colon (D, \mu|_D)\to (W_-, \omega_-)$ and $\theta_+\colon (R, \mu|_R)\to (W_+, \omega_+)$ denote the ergodic decompositions for $\mathcal{T}_-$ and $\mathcal{T}_+$, respectively.
Since we have $[\mathcal{T}:\mathcal{S}]_x<\infty$ for a.e.\ $x\in X$, the set $\tau^{-1}(w)$ is finite for a.e.\ $w\in W$ by Lemma \ref{lem-erg-dec}.
We define a Borel map $\psi \colon X\to \mathbb{R}_+^{\times}$ by $\psi(x)=\nu_{\theta(x)}(X_{\pi(x)})$ for $x\in X$.
Applying Lemma \ref{lem-rest} to $\theta$, $\pi$ and $\tau$, we have the equality $\nu_{\theta(x)}|_{X_{\pi(x)}}=\psi(x)\mu_{\pi(x)}$ for a.e.\ $x\in X$.
Combining the two equalities
\begin{align*}
U_*(\mu_{\pi(x)}|_{D_{\pi_-(x)}})&=\mathfrak{D}_{\mathcal{S}}(\phi, x)\mu_{\pi(U(x))}|_{R_{\pi_+(U(x))}},\\
U_*(\nu_{\theta(x)}|_{D_{\theta_-(x)}})&=\mathfrak{D}_{\mathcal{T}}(\phi, x)\nu_{\theta(U(x))}|_{R_{\theta_+(U(x))}}
\end{align*}
defining $\mathfrak{D}_{\mathcal{S}}(\phi, x)$ and $\mathfrak{D}_{\mathcal{T}}(\phi, x)$, we obtain the equality in the lemma.
\end{proof}
\begin{lem}\label{lem-d-cohom}
Let $A$ be a Borel subset of $X$ with $\mu(A)>0$.
We put $\mathfrak{D}=\mathfrak{D}(\mathcal{G}, \mathcal{S})$ and $\mathfrak{D}_A=\mathfrak{D}((\mathcal{G})_A, (\mathcal{S})_A)$.
Then there exists a Borel map $\psi \colon A\to \mathbb{R}_+^{\times}$ with the equality $\mathfrak{D}_A(g)=\psi(r(g))\mathfrak{D}(g)\psi(s(g))^{-1}$ for a.e.\ $g\in (\mathcal{G})_A$.
\end{lem}
\begin{proof}
We use the same notation as in the beginning of this subsection to define $\mathfrak{D}(\phi, x)$ for $x\in D$.
Fix $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$ with $\mu(D_{\phi})>0$.
Put $D=D_{\phi}$ and $R=R_{\phi}$.
We assume $D\subset A$ and $R\subset A$.
Put $\nu =\mu|_A$.
Let $\theta \colon (A, \nu)\to (W, \omega)$ be the ergodic decomposition for $(\mathcal{S})_A$ with $\theta_*\nu =\omega$.
Let $\nu=\int_W \nu_w d\omega(w)$ be the disintegration with respect to $\theta$.
To define the number $\mathfrak{D}_A(\phi, x)=\mathfrak{D}((\mathcal{G})_A, (\mathcal{S})_A, \phi, x)$, we use $\theta$ and $\{ \nu_w\}_w$ in place of $\pi$ and $\{ \mu_z\}_z$, and use the same $\mathcal{S}_-$, $\mathcal{S}_+$, $\pi_-$ and $\pi_+$ because we have $D\subset A$ and $R\subset A$.
For a.e.\ $x\in D$, we thus have the equality
\[U_*(\nu_{\theta(x)}|_{D_{\pi_-(x)}})=\mathfrak{D}_A(\phi, x)\nu_{\theta(U(x))}|_{R_{\pi_+(U(x))}}.\]
Let $Z_0$ be the Borel subset of $Z$ with $\pi^{-1}(Z_0)=\mathcal{S}A$.
Let $\alpha \colon W\to Z_0$ be the Borel isomorphism with $\pi =\alpha \circ \theta$ on $A$.
We have the disintegration
\[\nu=\int_W\mu_{\alpha(w)}(A\cap X_{\alpha(w)})^{-1}\mu_{\alpha(w)}|_A\, d\omega(w)\]
with respect to $\theta$ because the equality $\alpha_*\omega =\chi \xi|_{Z_0}$ holds, where $\chi$ is the Borel function on $Z_0$ defined by $\chi(z)=\mu_z(A\cap X_z)$ for $z\in Z_0$.
For a.e.\ $w\in W$, we have $\mu_{\alpha(w)}|_A=\mu_{\alpha(w)}(A\cap X_{\alpha(w)})\nu_w$ by uniqueness of disintegration.
We define a Borel map $\psi \colon A\to \mathbb{R}_+^{\times}$ by $\psi(x)=\mu_{\pi(x)}(A\cap X_{\pi(x)})$ for $x\in A$.
Combining the two equalities defining $\mathfrak{D}(\phi, x)$ and $\mathfrak{D}_A(\phi, x)$, we obtain the equality in the lemma.
\end{proof}
\subsection{The local-index cocycle}\label{subsec-lic}
We fix a standard finite measure space $(X, \mu)$ and a measure-preserving, discrete measured groupoid $\mathcal{G}$ on $(X, \mu)$.
Let $\mathcal{S}$ be a quasi-normal subgroupoid of $\mathcal{G}$.
We define a Borel function $\mathfrak{I}=\mathfrak{I}(\mathcal{G}, \mathcal{S})\colon \mathcal{G}\to \mathbb{R}_+^{\times}$ as follows.
Pick a countable family $\Phi$ of elements of ${\rm QN}_{\mathcal{G}}(\mathcal{S})$ such that $\mu(D_{\phi})>0$ for any $\phi \in \Phi$ and the equality $\mathcal{G}=\bigsqcup_{\phi \in \Phi}\phi(D_{\phi})$ holds.
For $\phi \in \Phi$, we put
\[\mathcal{S}_-=(\mathcal{S})_{D_{\phi}}\cap U_{\phi}^{-1}((\mathcal{S})_{R_{\phi}}),\quad \mathcal{S}_+=(\mathcal{S})_{R_{\phi}}\cap U_{\phi}((\mathcal{S})_{D_{\phi}}).\]
For $g\in \mathcal{G}$, choosing $\phi \in \Phi$ with $g\in \phi(D_{\phi})$, we define the number $\mathfrak{I}(g)\in \mathbb{R}_+^{\times}$ by
\begin{align*}
\mathfrak{I}(g)&=[[(\mathcal{S})_{R_{\phi}}:\mathcal{S}_+]]_{r(g)}[[(\mathcal{S})_{D_{\phi}}:\mathcal{S}_-]]_{s(g)}^{-1}\\
&=[[(\mathcal{S})_{R_{\phi}}:\mathcal{S}_+]]_{r(g)}[[U_{\phi}((\mathcal{S})_{D_{\phi}}):\mathcal{S}_+]]_{r(g)}^{-1}.
\end{align*}
By Lemma \ref{lem-li-res}, this definition of $\mathfrak{I}$ does not depend on the choice of $\Phi$.
We call $\mathfrak{I}$ the {\it local-index cocycle} for $\mathcal{G}$ and $\mathcal{S}$.
\begin{lem}
The function $\mathfrak{I} \colon \mathcal{G}\to \mathbb{R}_+^{\times}$ defined above is a Borel cocycle, that is, it preserves products.
\end{lem}
\begin{proof}
For each $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{S})$, we set $\mathcal{S}^{\phi}=U_{\phi}((\mathcal{S})_{D_{\phi}})$.
Pick two elements $\phi$, $\psi$ of ${\rm QN}_{\mathcal{G}}(\mathcal{S})$, and put $\eta =\psi \bullet \phi$, $D=D_{\eta}$ and $R=R_{\eta}$.
We assume $\mu(D)>0$ and the equalities $D=D_{\phi}$ and $R=R_{\psi}$.
For a.e.\ $x\in D$, putting $y=r\circ \eta(x)$, we obtain the equality
\[
\mathfrak{I}(\psi, r\circ \phi(x))=\frac{[[(\mathcal{S})_R :(\mathcal{S})_R\cap \mathcal{S}^{\psi}]]_y}{[[\mathcal{S}^{\psi}:(\mathcal{S})_R\cap \mathcal{S}^{\psi}]]_y}=\frac{[[(\mathcal{S})_R:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}{[[\mathcal{S}^{\psi}:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y},
\]
where we apply Lemma \ref{lem-li-prod} to the second equality.
Applying the same lemma, we also obtain the equality
\[
\mathfrak{I}(\phi, x)=\frac{[[\mathcal{S}^{\psi}:\mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}{[[\mathcal{S}^{\eta}:\mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}=\frac{[[\mathcal{S}^{\psi}:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}{[[\mathcal{S}^{\eta}:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}
\]
for a.e.\ $x\in D$ with $y=r\circ \eta(x)$.
Combining these two equalities, we have
\begin{align*}
\mathfrak{I}(\psi, r\circ \phi(x))\mathfrak{I}(\phi, x)&=\frac{[[(\mathcal{S})_R:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}{[[\mathcal{S}^{\eta}:(\mathcal{S})_R\cap \mathcal{S}^{\psi}\cap \mathcal{S}^{\eta}]]_y}=\frac{[[(\mathcal{S})_R:(\mathcal{S})_R\cap \mathcal{S}^{\eta}]]_y}{[[\mathcal{S}^{\eta}:(\mathcal{S})_R\cap \mathcal{S}^{\eta}]]_y}\\
&=\mathfrak{I}(\eta, x)
\end{align*}
for a.e.\ $x\in D$ with $y=r\circ \eta(x)$.
The function $\mathfrak{I}$ is therefore a Borel cocycle.
\end{proof}
\begin{lem}\label{lem-lic-res}
The following assertions hold:
\begin{enumerate}
\item Let $A$ be a Borel subset of $X$ with $\mu(A)>0$.
We put $\mathfrak{I}_A=\mathfrak{I}((\mathcal{G})_A, (\mathcal{S})_A)$.
Then the equality $\mathfrak{I}_A(g)=\mathfrak{I}(g)$ holds for a.e.\ $g\in (\mathcal{G})_A$.
\item Let $\mathcal{T}$ be a subgroupoid of $\mathcal{G}$ such that $\mathcal{S}<\mathcal{T}$; $[\mathcal{T}:\mathcal{S}]_x<\infty$ for a.e.\ $x\in X$; and $\mathcal{T}$ is quasi-normal in $\mathcal{G}$.
We set $\mathfrak{I}_{\mathcal{T}}=\mathfrak{I}(\mathcal{G}, \mathcal{T})$.
Then there exists a Borel map $\psi \colon X\to \mathbb{R}_+^{\times}$ with the equality $\mathfrak{I}_{\mathcal{T}}(g)=\psi(r(g))\mathfrak{I}(g)\psi(s(g))^{-1}$ for a.e.\ $g\in \mathcal{G}$.
\end{enumerate}
\end{lem}
\begin{proof}
Assertion (i) follows from Lemma \ref{lem-li-res}.
We prove assertion (ii).
We pick an element $\phi$ of the set ${\rm QN}_{\mathcal{G}}(\mathcal{T})={\rm QN}_{\mathcal{G}}(\mathcal{S})$ with $\mu(D_{\phi})>0$.
Put $D=D_{\phi}$ and $R=R_{\phi}$.
Define a Borel map $\psi \colon X\to \mathbb{R}_+^{\times}$ by $\psi(x)=[[\mathcal{T}: \mathcal{S}]]_x$ for $x\in X$.
For a.e.\ $x\in D$, putting $y=r\circ \phi(x)$, we obtain the equality
\begin{align*}
\mathfrak{I}_{\mathcal{T}}(\phi(x))&=\frac{[[(\mathcal{T})_R: (\mathcal{T})_R\cap \mathcal{T}^{\phi}]]_y}{[[\mathcal{T}^{\phi}: (\mathcal{T})_R\cap \mathcal{T}^{\phi}]]_y}=\frac{[[(\mathcal{T})_R: (\mathcal{S})_R\cap \mathcal{S}^{\phi}]]_y}{[[\mathcal{T}^{\phi}: (\mathcal{S})_R\cap \mathcal{S}^{\phi}]]_y}\\
&=\frac{[[(\mathcal{T})_R: (\mathcal{S})_R]]_y}{[[\mathcal{T}^{\phi}: \mathcal{S}^{\phi}]]_y}\frac{[[(\mathcal{S})_R: (\mathcal{S})_R\cap \mathcal{S}^{\phi}]]_y}{[[\mathcal{S}^{\phi}: (\mathcal{S})_R\cap \mathcal{S}^{\phi}]]_y}=\psi(y)\mathfrak{I}(\phi(x))\psi(x)^{-1},
\end{align*}
where we apply Lemma \ref{lem-li-prod} to the second and third equalities, and apply Lemma \ref{lem-li-res} to the fourth equality.
Assertion (ii) follows.
\end{proof}
\subsection{Computation}\label{subsec-comp}
For a finite-measure-preserving action of a Baumslag-Solitar group $\Gamma$, we show relationship between the modular homomorphism $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$ and the cocycles $\mathfrak{D}$ and $\mathfrak{I}$ for the groupoid associated with the action of $\Gamma$ and its elliptic subgroupoid.
\begin{lem}\label{lem-comp}
Let $\Gamma$ be a discrete group, and let $N$ be a quasi-normal subgroup of $\Gamma$.
Suppose that we have a measure-preserving action of $\Gamma$ on a standard finite measure space $(X, \mu)$.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{N}=N\ltimes X$, and set $\mathfrak{D}=\mathfrak{D}(\mathcal{G}, \mathcal{N})$ and $\mathfrak{I}=\mathfrak{I}(\mathcal{G}, \mathcal{N})$.
For any $\gamma \in \Gamma$, if both $N\cap \gamma N\gamma^{-1}$ and $N\cap \gamma^{-1}N\gamma$ are normal subgroups of $N$, then for a.e.\ $x\in X$, we have the equality
\[\mathfrak{D}(\gamma, x)\mathfrak{I}(\gamma, x)=[N: N\cap \gamma N\gamma^{-1}][N: N\cap \gamma^{-1}N\gamma]^{-1}.\]
\end{lem}
\begin{proof}
Fix $\gamma \in \Gamma$ and define $\phi \colon X\to \mathcal{G}$ by $\phi(x)=(\gamma, x)$ for $x\in X$.
We then have $\phi \in {\rm QN}_{\mathcal{G}}(\mathcal{N})$.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ be the ergodic decomposition for $\mathcal{N}$ with $\pi_*\mu=\xi$.
Let $\mu =\int_Z \mu_z d\xi(z)$ be the disintegration with respect to $\pi$.
We set $N_-=N\cap \gamma^{-1}N\gamma$ and $N_+=N\cap \gamma N\gamma^{-1}$.
Similarly, we set $\mathcal{N}_-=N_-\ltimes X$ and $\mathcal{N}_+=N_+\ltimes X$.
Let $\pi_-\colon (X, \mu)\to (Z_-, \xi_-)$ and $\pi_+\colon (X, \mu)\to (Z_+, \xi_+)$ be the ergodic decompositions for $\mathcal{N}_-$ and $\mathcal{N}_+$, respectively.
We have the canonical Borel maps $\sigma_-\colon (Z_-, \xi_-)\to (Z, \xi)$ and $\sigma_+\colon (Z_+, \xi_+)\to (Z, \xi)$ with $\pi =\sigma_-\circ \pi_-=\sigma_+\circ \pi_+$.
The group $N$ transitively acts on $\sigma_-^{-1}(\pi(x))$ and on $\sigma_+^{-1}(\pi(y))$ for a.e.\ $x\in X$ and a.e.\ $y\in X$ because $N_-$ and $N_+$ are normal subgroups of $N$.
By Lemma \ref{lem-li-group}, the equalities
\[[[\mathcal{N}:\mathcal{N}_-]]_x=\frac{[N:N_-]}{|\sigma_-^{-1}(\pi(x))|},\quad [[\mathcal{N}:\mathcal{N}_+]]_y=\frac{[N:N_+]}{|\sigma_+^{-1}(\pi(y))|}\]
hold for a.e.\ $x\in X$ and a.e.\ $y\in X$.
On the other hand, we have
\[\mu_{\pi(x)}(\pi_-^{-1}(\pi_-(x)))=|\sigma_-^{-1}(\pi(x))|^{-1},\quad \mu_{\pi(y)}(\pi_+^{-1}(\pi_+(y)))=|\sigma_+^{-1}(\pi(y))|^{-1}\]
for a.e.\ $x\in X$ and a.e.\ $y\in X$.
The equality $\mathfrak{D}(\gamma, x)\mathfrak{I}(\gamma, x)=[N: N_+][N:N_-]^{-1}$ thus holds for a.e.\ $x\in X$.
\end{proof}
\begin{cor}\label{cor-bs-comp}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$ and denote by $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$ the modular homomorphism.
Let $E$ be an infinite elliptic subgroup of $\Gamma$.
Suppose that we have a measure-preserving action of $\Gamma$ on a standard finite measure space $(X, \mu)$.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{E}=E\ltimes X$, and set $\mathfrak{D}=\mathfrak{D}(\mathcal{G}, \mathcal{E})$ and $\mathfrak{I}=\mathfrak{I}(\mathcal{G}, \mathcal{E})$.
Then we have the equality $\mathfrak{D}(\gamma, x)\mathfrak{I}(\gamma, x)=\mathfrak{m}(\gamma)$ for any $\gamma \in \Gamma$ and a.e.\ $x\in X$.
\end{cor}
\section{The Mackey range of the modular cocycle}\label{sec-mackey}
Fundamental ideas of the Mackey range of a cocycle are discussed in \cite[Section 6]{mackey} and \cite[Section 7]{ramsay}, where it is called the ``range closure'' of the cocycle.
Let $\mathcal{G}$ be a discrete measured groupoid on a standard finite measure space $(X, \mu)$.
Let $H$ be a locally compact second countable group and $\tau \colon \mathcal{G}\to H$ a Borel cocycle.
We define the Mackey range of $\tau$ as follows.
Define a discrete measured equivalence relation $\mathcal{R}$ on $X\times H$ by
\[\mathcal{R}=\{\, ((r(g), \tau(g)h), (s(g), h))\in (X\times H)^2\mid g\in \mathcal{G},\ h\in H\,\}.\]
Let $Z$ denote the space of ergodic components for $\mathcal{R}$.
The action of $H$ on $X\times H$ defined by $h(x, h')=(x, h'h^{-1})$ for $x\in X$ and $h, h'\in H$ induces an action of $H$ on $Z$.
This action $H\c Z$ is called the {\it Mackey range} of $\tau$.
We can check the following two lemmas that are basic facts on Mackey ranges.
\begin{lem}\label{lem-mackey-res}
In the above notation, let $A$ be a Borel subset of $X$ with $\mathcal{G}A=X$.
We denote by $H\c Z_A$ the Mackey range of the restriction of $\tau$ to $(\mathcal{G})_A$.
Then the inclusion of $A\times H$ into $X\times H$ induces an isomorphism between the two actions $H\c Z_A$ and $H\c Z$.
Namely, it induces a measure space isomorphism $\eta \colon Z_A\to Z$ such that for any $h\in H$, we have $\eta(hz)=h\eta(z)$ for a.e.\ $z\in Z_A$.
\end{lem}
\begin{lem}\label{lem-mackey-cohom}
In the above notation, the following assertions hold:
\begin{enumerate}
\item Let $f$ be a Borel automorphism of $\mathcal{G}$.
Define a Borel cocycle $\tau_f\colon \mathcal{G}\to H$ by $\tau_f=\tau\circ f$.
Then the automorphism of $X\times H$ sending $(x, h)$ to $(f^{-1}(x), h)$ for $x\in X$ and $h\in H$ induces an isomorphism between the Mackey ranges of $\tau$ and $\tau_f$.
\item Let $\varphi \colon X\to H$ be a Borel map.
Define a Borel cocycle $\tau_{\varphi}\colon \mathcal{G}\to H$ by $\tau_{\varphi}(g)=\varphi(r(g))\tau(g)\varphi(s(g))^{-1}$ for $g\in \mathcal{G}$.
Then the automorphism of $X\times H$ sending $(x, h)$ to $(x, \varphi(x)h)$ for $x\in X$ and $h\in H$ induces an isomorphism between the Mackey ranges of $\tau$ and $\tau_{\varphi}$.
\end{enumerate}
\end{lem}
In the rest of this section, we set $\Gamma ={\rm BS}(p, q)$ and $\Lambda ={\rm BS}(r, s)$ with $2\leq |p|\leq |q|$ and $2\leq |r|\leq |s|$, unless otherwise mentioned.
Let $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$ and $\mathfrak{n} \colon \Lambda \to \mathbb{R}_+^{\times}$ denote the modular homomorphisms.
\begin{thm}\label{thm-r}
Let $(\Sigma, m)$ be a $(\Gamma, \Lambda)$-coupling.
Then there exists an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to (\mathbb{R}, \log \circ \mathfrak{m}, \log \circ \mathfrak{n})$.
\end{thm}
\begin{proof}
We may assume that the action of $\Gamma \times \Lambda$ on $(\Sigma, m)$ is ergodic by \cite[Corollary 3.6]{fmw}.
Let $X, Y\subset \Sigma$ be fundamental domains for the actions of $\Lambda$ and $\Gamma$ on $\Sigma$, respectively.
Putting $Z=X\cap Y$, we may assume that $Z$ has positive measure.
Let $\alpha \colon \Gamma \times X\to \Lambda$ be the ME cocycle associated to $X$, defined by the condition $(\gamma, \alpha(\gamma, x))x\in X$ for $\gamma \in \Gamma$ and $x\in X$.
Since $X$ is identified with $\Sigma /(\{ e\} \times \Lambda)$, we have an ergodic measure-preserving action $\Gamma \c X$, where $X$ is equipped with the finite measure that is the restriction of $m$.
To distinguish this action from the original action on $\Sigma$, we use a dot for the new action, that is, we denote $(\gamma, \alpha(\gamma, x))x$ by $\gamma \cdot x$.
Similarly, we have an ergodic measure-preserving action $\Lambda \c Y$ and use a dot for this action.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{H}=\Lambda \ltimes Y$.
The map $f\colon (\mathcal{G})_Z\to (\mathcal{H})_Z$ defined by $f(\gamma, x)=(\alpha(\gamma, x), x)$ for $(\gamma, x)\in (\mathcal{G})_Z$ is then an isomorphism.
Let $T_{\Gamma}$ and $T_{\Lambda}$ denote the Bass-Serre trees associated to $\Gamma$ and $\Lambda$, respectively.
For a vertex $u$ of $T_{\Gamma}$, we set $\mathcal{G}_u=\Gamma_u\ltimes X$.
Similarly, for a vertex $v$ of $T_{\Lambda}$, we set $\mathcal{H}_v=\Lambda_v\ltimes Y$.
We fix a vertex $u$ of $T_{\Gamma}$.
By Theorem \ref{thm-ell}, $f((\mathcal{G}_u)_Z)$ is an elliptic subgroupoid of $(\mathcal{H})_Z$.
It follows that there exist a Borel subset $Z_1$ of $Z$ with positive measure and a vertex $v$ of $T_{\Lambda}$ with $f((\mathcal{G}_u)_{Z_1})<(\mathcal{H}_v)_{Z_1}$.
Repeating this argument for $f^{-1}$ and $\mathcal{H}_v$, we find a Borel subset $A$ of $Z_1$ with positive measure and a vertex $u'$ of $T_{\Gamma}$ with $(\mathcal{H}_v)_A<f((\mathcal{G}_{u'})_A)$.
Putting $\mathcal{E}=f^{-1}((\mathcal{H}_v)_A)$, we have
\[((\Gamma_u\cap \Gamma_{u'})\ltimes X)_A<\mathcal{E}<(\Gamma_{u'}\ltimes X)_A.\]
We put $\mathfrak{D}_{\Gamma}=\mathfrak{D}((\mathcal{G})_A, \mathcal{E})$ and $\mathfrak{I}_{\Gamma}=\mathfrak{I}((\mathcal{G})_A, \mathcal{E})$.
Let $\rho \colon \mathcal{G}\to \Gamma$ be the projection.
Since $\Gamma_u\cap \Gamma_{u'}$ is a finite index subgroup of $\Gamma_{u'}$, by Lemmas \ref{lem-d-finite}, \ref{lem-d-cohom} and \ref{lem-lic-res} and Corollary \ref{cor-bs-comp}, there exists a Borel map $\varphi_0\colon A\to \mathbb{R}_+^{\times}$ such that
\[\mathfrak{D}_{\Gamma}(g)\mathfrak{I}_{\Gamma}(g)=\varphi_0(r_{\mathcal{G}}(g))\mathfrak{m} \circ \rho(g)\varphi_0(s_{\mathcal{G}}(g))^{-1}\]
for a.e.\ $g\in (\mathcal{G})_A$, where $r_{\mathcal{G}}, s_{\mathcal{G}}\colon \mathcal{G}\to X$ denote the range and source maps of $\mathcal{G}$, respectively.
Put $\mathfrak{D}_{\Lambda}=\mathfrak{D}((\mathcal{H})_A, (\mathcal{H}_v)_A)$ and $\mathfrak{I}_{\Lambda}=\mathfrak{I}((\mathcal{H})_A, (\mathcal{H}_v)_A)$.
Let $\sigma \colon \mathcal{H}\to \Lambda$ be the projection.
A similar property also holds for these maps.
On the other hand, the equality $f(\mathcal{E})=(\mathcal{H}_v)_A$ implies that $\mathfrak{D}_{\Lambda}\circ f=\mathfrak{D}_{\Gamma}$ and $\mathfrak{I}_{\Lambda}\circ f=\mathfrak{I}_{\Gamma}$.
It follows that there exists a Borel map $\varphi \colon A\to \mathbb{R}_+^{\times}$ such that
\[\varphi(r_{\mathcal{G}}(g))\mathfrak{m} \circ \rho(g)\varphi(s_{\mathcal{G}}(g))^{-1}=\mathfrak{n} \circ \sigma\circ f(g)\]
for a.e.\ $g\in (\mathcal{G})_A$.
We thus have the equality
\[\varphi(\gamma \cdot x)\mathfrak{m}(\gamma)\varphi(x)^{-1}=\mathfrak{n} \circ \alpha(\gamma, x)\tag{$\ast$}\label{eq-cohom}\]
for any $\gamma \in \Gamma$ and a.e.\ $x\in A$ with $\gamma \cdot x\in A$.
We define a Borel map $\Phi \colon \Sigma \to \mathbb{R}$ by
\[\Phi((\gamma, \lambda)x)=\log (\mathfrak{m}(\gamma)\varphi(x)^{-1}\mathfrak{n}(\lambda)^{-1})\]
for $x\in A$, $\gamma \in \Gamma$ and $\lambda \in \Lambda$.
Using equation (\ref{eq-cohom}), we can show that $\Phi$ is well-defined and is a desired map, along the argument in the proof of \cite[Theorem 4.4]{kida-ama}.
\end{proof}
Theorem \ref{thm-flow} is obtained by combining Lemma \ref{lem-mackey-res}, Lemma \ref{lem-mackey-cohom} and equation (\ref{eq-cohom}).
\begin{rem}\label{rem-type}
Let $\Gamma \c (X, \mu)$ be an ergodic measure-preserving action on a standard finite measure space, and call it $\alpha$.
We denote by $\pi \colon (X, \mu)\to (Z, \xi)$ the ergodic decomposition for the action of $\ker \mathfrak{m}$ on $(X, \mu)$.
We have the canonical action of $\mathfrak{m}(\Gamma)$ on $(Z, \xi)$ and the action of $\log \circ \mathfrak{m}(\Gamma)$ on it through the isomorphism $\log \colon \mathbb{R}_+^{\times} \to \mathbb{R}$.
The flow associated with $\alpha$, introduced in Section \ref{sec-int}, is defined as the action of $\mathbb{R}$ induced from the action of $\log \circ \mathfrak{m}(\Gamma)$ on $(Z, \xi)$.
If $|p|=|q|$, then the flow associated with $\alpha$ is isomorphic to the action of $\mathbb{R}$ on itself by addition because $\mathfrak{m}$ is trivial.
If $|p|<|q|$, then we have the following two cases:
\begin{itemize}
\item If there exists a positive integer $n$ such that up to null sets, $Z$ consists of $n$ points each of whose measure is $\xi(Z)/n$, then the flow associated with $\alpha$ is isomorphic to the action of $\mathbb{R}$ on $\mathbb{R}/(-\log |p/q|^n)\mathbb{Z}$ by addition.
\item If $Z$ contains no point whose measure is positive, then the flow associated with $\alpha$ is finite-measure-preserving and any orbit of it is of measure zero.
\end{itemize}
Following terminology for transformations of type ${\rm III}$ (see Section \ref{sec-exotic}), in the first case, we say that the action $\Gamma \c (X, \mu)$ is of {\it type $|p/q|^n$}.
In the second case, we say that the action $\Gamma \c (X, \mu)$ is of {\it type $0$}.
\end{rem}
\begin{cor}
Assume $|p|<|q|$ and $|r|<|s|$.
If $\Gamma$ and $\Lambda$ are ME, then $\ker \mathfrak{m}$ and $\ker \mathfrak{n}$ are ME.
\end{cor}
\begin{proof}
We may assume $|q/p|\leq |s/r|$.
Let $\Sigma$ be a $(\Gamma, \Lambda)$-coupling.
There exists an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to (\mathbb{R}, \log \circ \mathfrak{m}, \log \circ \mathfrak{n})$ by Theorem \ref{thm-r}.
We set $U=\{\, u\in \mathbb{R}\mid 0\leq u<\log |q/p|\,\}$, which is a fundamental domain for the action of $\log \circ \mathfrak{m}(\Gamma)$ on $\mathbb{R}$ and is contained in that for the action of $\log \circ \mathfrak{n}(\Lambda)$ on $\mathbb{R}$.
The set $\Phi^{-1}(U)$ then has positive measure and is a $(\ker \mathfrak{m}, \ker \mathfrak{n})$-coupling.
\end{proof}
For discrete groups $G$ and $H$, two measure-preserving actions $G \c (Z, \xi)$ and $H \c (W, \omega)$ on measure spaces are called {\it conjugate} if there exist an isomorphism $F\colon G\to H$ and a measurable isomorphism $f$ from a conull measurable subset of $Z$ onto a conull measurable subset of $W$ such that $f_*\xi$ and $\omega$ are equivalent and for any $g\in G$, we have $f(gz)=F(g)f(z)$ for a.e.\ $z\in Z$.
\begin{cor}\label{cor-conj}
Assume $|p/q|=|r/s|$.
Let $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ be ergodic f.f.m.p.\ actions.
Let $\pi \colon (X, \mu)\to (Z, \xi)$ and $\theta \colon (Y, \nu)\to (W, \omega)$ be the ergodic decompositions for the actions $\ker \mathfrak{m} \c (X, \mu)$ and $\ker \mathfrak{n} \c (Y, \nu)$, respectively.
If the actions $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ are WOE, then the canonical actions $\mathfrak{m}(\Gamma)\c (Z, \xi)$ and $\mathfrak{n}(\Lambda)\c (W, \omega)$ are conjugate.
\end{cor}
\begin{proof}
We have the actions $\log \circ \mathfrak{m}(\Gamma)\c (Z, \xi)$ and $\log \circ \mathfrak{n}(\Lambda)\c (W, \omega)$ through the isomorphism $\log \colon \mathbb{R}_+^{\times}\to \mathbb{R}$.
Let $\mathbb{R} \c (Z_1, \xi_1)$ and $\mathbb{R} \c (W_1, \omega_1)$ denote the actions induced from these two actions, respectively.
By Theorem \ref{thm-flow}, these two actions of $\mathbb{R}$ are isomorphic.
The assumption $|p/q|=|r/s|$ implies the equality $\log \circ \mathfrak{m}(\Gamma)=\log \circ \mathfrak{n}(\Lambda)$.
The action of $\log \circ \mathfrak{m}(\Gamma)$ on almost every ergodic component for its action on $(Z_1, \xi_1)$ is isomorphic to the original action $\log \circ \mathfrak{m}(\Gamma)\c (Z, \xi)$.
A similar property holds for the actions of $\log \circ \mathfrak{n}(\Lambda)$ on $(W_1, \omega_1)$ and on $(W, \omega)$.
The corollary therefore follows.
\end{proof}
\begin{rem}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$, and suppose that we have an ergodic measure-preserving action of $\mathfrak{m}(\Gamma)$ on a standard finite measure space $(Z, \xi)$.
We construct an ergodic f.f.m.p.\ action $\Gamma \c (X, \mu)$ such that the canonical action of $\mathfrak{m}(\Gamma)$ on the space of ergodic components for the action $\ker \mathfrak{m} \c (X, \mu)$ is isomorphic to the given action on $(Z, \xi)$.
Let $\Gamma \c (Y, \nu)$ be an f.f.m.p.\ action such that its restriction to $\ker \mathfrak{m}$ is ergodic.
We set $(X, \mu)=(Y, \nu)\times (Z, \xi)$.
Define an action of $\Gamma$ on $(X, \mu)$ by $\gamma(y, z)=(\gamma y, \mathfrak{m}(\gamma)z)$ for $\gamma \in \Gamma$, $y\in Y$ and $z\in Z$.
This is shown to be a desired action.
By definition, any associated flow is induced from an ergodic measure-preserving action of $\log \circ \mathfrak{m}(\Gamma)$ on a standard finite measure space.
Thanks to the construction discussed above, we can find an ergodic f.f.m.p.\ action of $\Gamma$ whose associated flow is isomorphic to such a given action of $\mathbb{R}$.
\end{rem}
\begin{proof}[Proof of Theorem \ref{thm-z}]
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|<|q|$, and suppose that an ergodic f.f.m.p.\ action $\Gamma \c (X, \mu)$ is WOE to a weakly mixing f.f.m.p.\ action $\Delta \c (Y, \nu)$ of a discrete group $\Delta$.
Let $\Sigma$ be the $(\Gamma, \Delta)$-coupling associated with the WOE between the actions $\Gamma \c (X, \mu)$ and $\Delta \c (Y, \nu)$.
Combining Theorems \ref{thm-abel} and \ref{thm-r}, we obtain a homomorphism $\delta \colon \Delta \to \mathbb{R}$ and an almost $(\Gamma \times \Delta)$-equivariant Borel map $\Phi \colon \Sigma \to (\mathbb{R}, \log \circ \mathfrak{m}, \delta)$.
If $\delta$ were trivial, then $\Phi$ would induce a $\Gamma$-equivariant Borel map from $X=\Sigma /\Delta$ into $\mathbb{R}$.
The group $\log \circ \mathfrak{m}(\Gamma)$ is infinite and discrete in $\mathbb{R}$.
This is a contradiction because we have the $\Gamma$-invariant finite measure $\mu$ on $X$.
The homomorphism $\delta$ is therefore non-trivial.
Put $M=(\log |q/p|)\mathbb{Z}$.
The map $\Phi$ induces a $\Delta$-equivariant Borel map $\bar{\Phi}$ from $Y=\Sigma /\Gamma$ into $\mathbb{R} /M$.
Define a Borel map $\Psi \colon Y\times Y\to \mathbb{R}/M$ by $\Psi(x, y)=\bar{\Phi}(x)-\bar{\Phi}(y)$ for $x, y\in Y$.
Let $\Delta$ act on $Y\times Y$ diagonally.
The map $\Psi$ is $\Delta$-invariant and is therefore essentially constant because the action $\Delta \c (Y, \nu)$ is weakly mixing.
It follows that $\bar{\Phi}$ is essentially constant and that $\delta(\Delta)$ is contained in $M$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm-bs-hyp}]
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|<|q|$.
Let $\Delta$ be a discrete group such that there is an infinite amenable normal subgroup $N$ of $\Delta$ with the quotient $\Delta /N$ non-elementarily word-hyperbolic.
Assuming that $\Gamma$ and $\Delta$ are ME, we deduce a contradiction.
Let $\Sigma$ be a $(\Gamma, \Delta)$-coupling such that the action of $\Gamma \times \Delta$ on $\Sigma$ is ergodic.
As in the proof of Theorem \ref{thm-r}, we choose fundamental domains $X, Y\subset \Sigma$ for the actions of $\Delta$ and $\Gamma$ on $\Sigma$, respectively, such that the intersection $Z=X\cap Y$ has positive measure.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{H}=\Delta \ltimes Y$.
Let $\alpha \colon \Gamma \times X\to \Delta$ be the ME cocycle associated to $X$.
We then have the isomorphism $f\colon (\mathcal{G})_Z\to (\mathcal{H})_Z$ defined by $f(\gamma, x)=(\alpha(\gamma, x), x)$ for $(\gamma, x)\in (\mathcal{G})_Z$.
Let $T$ denote the Bass-Serre tree associated to $\Gamma$.
For a vertex $v$ of $T$, set $\mathcal{G}_v=\Gamma_v\ltimes X$.
Since $(\mathcal{G}_v)_Z$ is amenable and quasi-normal in $(\mathcal{G})_Z$, by Theorem \ref{thm-ell-hyp}, there exist a Borel subset $Z_1$ of $Z$ with positive measure and a subgroup $L$ of $\Delta$ with $N<L$, $[L: N]<\infty$ and $f((\mathcal{G}_v)_{Z_1})<(L\ltimes X)_{Z_1}$.
Since $(L\ltimes X)_{Z_1}$ is also amenable and quasi-normal in $(\mathcal{H})_{Z_1}$, by Theorem \ref{thm-ell}, there exist a Borel subset $A$ of $Z_1$ with positive measure and a vertex $v'$ of $T$ with $f^{-1}((L\ltimes X)_A)<(\mathcal{G}_{v'})_A$.
Along the proof of Theorem \ref{thm-r}, using Lemma \ref{lem-comp}, we find a $(\Gamma \times \Delta)$-equivariant Borel map $\Phi \colon \Sigma \to (\mathbb{R}, \log \circ \mathfrak{m}, \log \circ I)$, where $I\colon \Delta \to \mathbb{R}_+^{\times}$ is the trivial homomorphism.
The map $\Phi$ induces a $\Gamma$-equivariant Borel map from $\Sigma /\Delta$ into $\mathbb{R}$.
This is a contradiction because $\log \circ \mathfrak{m}(\Gamma)$ is infinite and discrete in $\mathbb{R}$.
\end{proof}
\section{Reduction of a coupling of Baumslag-Solitar groups}\label{sec-red}
For a discrete group $G$ acting on a tree $T$ so that the stabilizer of any simplex of $T$ in $G$ is infinite and cyclic, Gilbert, Howie, Metaftsis and Raptis present a sufficient condition for the automorphism group of $G$ to be canonically represented into the automorphism group of $T$, in \cite[Theorem A]{ghmr}.
In Section \ref{subsec-red}, relying on their argument, for any coupling of two Baumslag-Solitar groups satisfying a certain ergodicity condition, we construct an equivariant Borel map from the coupling into the space of isomorphisms between the associated Bass-Serre trees.
Throughout this section, we fix the following:
\begin{notation}\label{not-bs}
Let $p$, $q$, $r$ and $s$ be integers with $2\leq |p|<|q|$ and $2\leq |r|<|s|$.
Let $d_0>0$ denote the greatest common divisor of $p$ and $q$, and let $c_0>0$ denote the greatest common divisor of $r$ and $s$.
Putting $p_0=p/d_0$, $q_0=q/d_0$, $r_0=r/c_0$ and $s_0=s/c_0$, we assume that $q$ is not a multiple of $p$ and that $s$ is not a multiple of $r$.
We thus have $1<|p_0|<|q_0|$ and $1<|r_0|<|s_0|$.
We set $\Gamma ={\rm BS}(p, q)$ and $\Lambda ={\rm BS}(r, s)$.
Let $T_{\Gamma}$ and $T_{\Lambda}$ be the Bass-Serre trees associated with $\Gamma$ and $\Lambda$, respectively.
We denote by $\imath \colon \Gamma \to {\rm Aut}(T_{\Gamma})$ and $\jmath \colon \Lambda \to {\rm Aut}(T_{\Lambda})$ the homomorphisms associated with the actions on the trees.
Let $a\in \Gamma$ (resp.\ $b\in \Lambda$) be a generator of the stabilizer of a vertex in $T_{\Gamma}$ (resp.\ $T_{\Lambda}$).
We define ${\rm Isom}(T_{\Lambda}, T_{\Gamma})$ as the set of orientation-preserving isomorphisms from $T_{\Lambda}$ onto $T_{\Gamma}$, which are equipped with the standard Borel structure induced by the pointwise convergence topology.
The set ${\rm Isom}(T_{\Gamma}, T_{\Lambda})$ is also defined similarly.
\end{notation}
\subsection{Reduction to isomorphisms between trees}\label{subsec-red}
For non-zero integers $m$, $n$, we say that a measure-preserving action of $\mathbb{Z}$ on a measure space is {\it $(m, n)$-ergodic} if for any non-negative integers $k$, $l$, the action of the subgroup $m^kn^l\mathbb{Z}$ is ergodic.
\begin{thm}\label{thm-aut-t}
In Notation \ref{not-bs}, let $(\Sigma, m)$ be a $(\Gamma, \Lambda)$-coupling.
We suppose that
\begin{enumerate}
\item[(1)] the action of $\langle a^{d_0}\rangle$ on almost every ergodic component for the action $\langle a\rangle \c \Sigma /\Lambda$ is $(p_0, q_0)$-ergodic; and
\item[(2)] the action of $\langle b^{c_0}\rangle$ on almost every ergodic component for the action $\langle b\rangle \c \Sigma /\Gamma$ is $(r_0, s_0)$-ergodic.
\end{enumerate}
Then there exists an essentially unique, almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to ({\rm Isom}(T_{\Lambda}, T_{\Gamma}), \imath, \jmath)$.
In particular, we have $|p|=|r|$ and $|q|=|s|$.
\end{thm}
\begin{lem}\label{lem-dirac}
In Notation \ref{not-bs}, the Dirac measure on the neutral element of ${\rm Aut}(T_{\Gamma})$ is the only probability measure on ${\rm Aut}(T_{\Gamma})$ invariant under conjugation by any element of $\imath(\Gamma)$.
\end{lem}
\begin{proof}
Let $\mu$ be a probability measure on ${\rm Aut}(T_{\Gamma})$ invariant under conjugation by any element of $\imath(\Gamma)$.
Let $\Gamma =\langle \, a, t\mid ta^pt^{-1}=a^q\,\rangle$ be the presentation of $\Gamma$.
Note that $t$ acts on $T_{\Gamma}$ as a hyperbolic isometry.
Let $x_{\pm}\in \partial T_{\Gamma}$ denote the two fixed points of $t$.
We define a Borel map $F\colon {\rm Aut}(T_{\Gamma})\to \partial T_{\Gamma}$ by $F(\varphi)=\varphi(x_+)$ for $\varphi \in {\rm Aut}(T_{\Gamma})$.
Since $tx_+=x_+$, we have $F_*\mu=(F\circ {\rm Ad} t)_*\mu=(\imath(t)\circ F)_*\mu$.
The measure $F_*\mu$ is thus invariant under the action of $t$.
Since the action of $\langle t\rangle$ on $\partial T_{\Gamma}\setminus \{ x_{\pm}\}$ is free and admits a Borel fundamental domain, we have $F_*\mu(\{ x_{\pm}\})=1$.
It follows that $\mu$ is supported on the set $\{\, \varphi \in {\rm Aut}(T_{\Gamma})\mid \varphi(x_+)\in \{ x_{\pm}\}\,\}$.
A verbatim argument implies that $\mu$ is supported on the stabilizer of the set $\{ x_{\pm}\}$ in ${\rm Aut}(T_{\Gamma})$.
The element $ata^{-1}\in \Gamma$ also acts on $T_{\Gamma}$ as a hyperbolic isometry.
In a similar way, if $y_{\pm}\in \partial T_{\Gamma}$ denote the two fixed points of $ata^{-1}$, then $\mu$ is supported on the stabilizer of the set $\{ y_{\pm}\}$ in ${\rm Aut}(T_{\Gamma})$.
Let $v$ denote the vertex of $T_{\Gamma}$ such that the stabilizer of $v$ in $\Gamma$ is equal to $\langle a\rangle$.
Any element $\varphi$ of ${\rm Aut}(T_{\Gamma})$ with $\varphi(\{ x_{\pm}\})=\{ x_{\pm}\}$ and $\varphi(\{ y_{\pm}\})=\{ y_{\pm}\}$ fixes $v$ because $v$ is the intersection of the geodesic in $T_{\Gamma}$ between $x_+$ and $x_-$ and the one between $y_+$ and $y_-$.
It follows that $\mu$ is supported on the stabilizer of $v$ in ${\rm Aut}(T_{\Gamma})$.
Since $\Gamma$ acts on $V(T_{\Gamma})$ transitively, we conclude that $\mu$ is supported on the neutral element of ${\rm Aut}(T_{\Gamma})$.
\end{proof}
In the notation in Theorem \ref{thm-aut-t}, let $X, Y\subset \Sigma$ be fundamental domains for the actions of $\Lambda$ and $\Gamma$ on $\Sigma$, respectively.
Let $\alpha \colon \Gamma \times X\to \Lambda$ be the ME cocycle defined by the condition $(\gamma, \alpha(\gamma, x))x\in X$ for $\gamma \in \Gamma$ and $x\in X$.
To distinguish the action $\Gamma \c X$ and the action $\Gamma \c \Sigma$, we use a dot for the former action, that is, we set $\gamma \cdot x=(\gamma, \alpha(\gamma, x))x$ for $\gamma \in \Gamma$ and $x\in X$.
Similarly, we use a dot for the action $\Lambda \c Y$.
By \cite[Lemma 2.27]{kida-survey}, we may assume that the intersection $Z=X\cap Y$ satisfies the equality $\Gamma \cdot Z=X$ when $Z$ is regarded as a subset of $X$, and satisfies the equality $\Lambda \cdot Z=Y$ when $Z$ is regarded as a subset of $Y$.
We set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{H}=\Lambda \ltimes Y$.
Let $f\colon (\mathcal{G})_Z\rightarrow (\mathcal{H})_Z$ be the isomorphism defined by $f(\gamma, x)=(\alpha(\gamma, x), x)$ for $(\gamma, x)\in (\mathcal{G})_Z$.
For each simplex $u$ of $T_{\Gamma}$, we put $\mathcal{G}_u=\Gamma_u\ltimes X$.
Similarly, for each simplex $v$ of $T_{\Lambda}$, we put $\mathcal{H}_v=\Lambda_v\ltimes Y$.
\begin{lem}\label{lem-vertex-stab}
If we have $u_1, u_2\in V(T_{\Gamma})$ and a Borel subset $W$ of $Z$ with positive measure satisfying the inclusion $(\mathcal{G}_{u_1})_W\subset (\mathcal{G}_{u_2})_W$, then the equality $u_1=u_2$ holds.
\end{lem}
\begin{proof}
Assuming that $u_1\neq u_2$, we deduce a contradiction.
Let $e$ be the edge of $T_{\Gamma}$ containing $u_1$ and contained in the geodesic between $u_1$ and $u_2$.
The inclusion in the lemma implies the equality $(\mathcal{G}_e)_W=(\mathcal{G}_{u_1})_W$.
On the other hand, by condition (1) in Theorem \ref{thm-aut-t} and Lemma \ref{lem-index} (ii), (iii), we have
$[(\mathcal{G}_{u_1})_W: (\mathcal{G}_e)_W]_x=[\Gamma_{u_1}: \Gamma_e]>1$ for a.e.\ $x\in W$.
This is a contradiction.
\end{proof}
\begin{lem}\label{lem-pre}
For any $u\in V(T_{\Gamma})$, there exist a countable Borel partition $Z=\bigsqcup_n Z_n$ and $v_n\in V(T_{\Lambda})$ such that $f((\mathcal{G}_u)_{Z_n})=(\mathcal{H}_{v_n})_{Z_n}$ for each $n$.
\end{lem}
\begin{proof}
By Lemmas \ref{lem-qn-group} and \ref{lem-qn-res}, $(\mathcal{G}_u)_Z$ is quasi-normal in $(\mathcal{G})_Z$, and it is also of infinite type and amenable.
It follows that $f((\mathcal{G}_u)_Z)$ is of infinite type, amenable and quasi-normal in $(\mathcal{H})_Z$.
Let $\sigma \colon \mathcal{H}\to \Lambda$ be the projection.
By Theorem \ref{thm-ell}, $f((\mathcal{G}_u)_Z)$ is elliptic, that is, there exists an $(f((\mathcal{G}_u)_Z), \sigma)$-invariant Borel map $\varphi \colon Z\to V(T_{\Lambda})$.
Take a countable Borel partition $Z=\bigsqcup_n Z_n$ such that for each $n$, the map $\varphi$ is essentially constant on $Z_n$, and denote by $v_n\in V(T_{\Lambda})$ the essential value of $\varphi$ on $Z_n$.
We obtain the inclusion $f((\mathcal{G}_u)_{Z_n})\subset (\mathcal{H}_{v_n})_{Z_n}$ for any $n$.
Applying the same argument to $f^{-1}$ in place of $f$ and taking a finer Borel partition of $Z$, for each $n$, we find $u_n\in V(T_{\Gamma})$ with the inclusion $f^{-1}((\mathcal{H}_{v_n})_{Z_n})\subset (\mathcal{G}_{u_n})_{Z_n}$.
By Lemma \ref{lem-vertex-stab}, for any $n$, the equality $u_n=u$ holds, and the equality in the lemma also holds.
\end{proof}
\begin{lem}\label{lem-edge-stab}
If we have $e_1, e_2\in E(T_{\Lambda})$ and a Borel subset $W$ of $Z$ with positive measure such that $(\mathcal{H}_{e_1})_W=(\mathcal{H}_{e_2})_W$, then the equality $\Lambda_{e_1}=\Lambda_{e_2}$ holds.
\end{lem}
\begin{proof}
Since we have $(\mathcal{H}_{e_1})_W=(\mathcal{H}_{e_1}\cap \mathcal{H}_{e_2})_W=(\mathcal{H}_{e_2})_W$, condition (2) in Theorem \ref{thm-aut-t} and Lemma \ref{lem-index} (ii), (iii) imply that we have $\Lambda_{e_1}=\Lambda_{e_1}\cap \Lambda_{e_2}=\Lambda_{e_2}$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{thm-aut-t}]
For a.e.\ $z\in Z$, we define a map $\varphi_z\colon V(T_{\Gamma})\to V(T_{\Lambda})$ by $\varphi_z(u)=v_n$ if $z\in Z_n$, where we use the notation in Lemma \ref{lem-pre}.
By symmetry, a conclusion similar to Lemma \ref{lem-vertex-stab} holds for subgroupoids of $\mathcal{H}$.
For a.e.\ $z\in Z$, the map $\varphi_z$ is thus well-defined and is independent of the choice of a countable Borel partition of $Z$.
Applying the same process to $f^{-1}$ in place of $f$, for a.e.\ $z\in Z$, we obtain a map $\psi_z\colon V(T_{\Lambda})\to V(T_{\Gamma})$.
By Lemma \ref{lem-vertex-stab}, for a.e.\ $z\in Z$, the composition $\psi_z\circ \varphi_z$ is the identity on $V(T_{\Gamma})$, and $\varphi_z\circ \psi_z$ is the identity on $V(T_{\Lambda})$.
In particular, $\varphi_z$ and $\psi_z$ are bijections.
\begin{claim}
For a.e.\ $z\in Z$, $\varphi_z$ defines a simplicial isomorphism from $T_{\Gamma}$ onto $T_{\Lambda}$.
\end{claim}
\begin{proof}
This proof heavily depends on the proof of \cite[Theorem A]{ghmr}.
We extend the map $\varphi_z\colon V(T_{\Gamma})\to V(T_{\Lambda})$ to a continuous map $\tilde{\varphi}_z\colon T_{\Gamma}\to T_{\Lambda}$ in an affine way.
Namely, for each edge $e$ of $T_{\Gamma}$ whose origin and terminal are $u_1$ and $u_2$, respectively, the restriction of $\tilde{\varphi}_z$ to $e$ is the affine homeomorphism from $e$ to the geodesic from $\varphi_z(u_1)$ to $\varphi_z(u_2)$.
Similarly, we extend $\psi_z$ to a continuous map $\tilde{\psi}_z\colon T_{\Lambda}\to T_{\Gamma}$.
We show that for a.e.\ $z\in Z$, $\tilde{\varphi}_z$ is locally injective, that is, for any $u\in V(T_{\Gamma})$ and any two distinct edges $e_1$, $e_2$ of $T_{\Gamma}$ whose boundaries contain $u$, the first edges of the paths $\tilde{\varphi}_z(e_1)$, $\tilde{\varphi}_z(e_2)$ are distinct.
This implies injectivity of $\tilde{\varphi}_z$, and thus implies the claim.
We pick $u\in V(T_{\Gamma})$ and suppose that we have a Borel subset $W$ of $Z$ with positive measure and two edges $e$, $e'$ of $T_{\Gamma}$ whose boundaries contain $u$, such that for a.e.\ $z\in W$, the first edges of the two paths $\tilde{\psi}_z\circ \tilde{\varphi}_z(e)$, $\tilde{\psi}_z\circ \tilde{\varphi}_z(e')$ in $T_{\Gamma}$ starting at $u$ are equal.
Denote the first edge by $d$.
It suffices to show the equality $e=e'$.
For a.e.\ $z\in Z$, any $\gamma \in \Gamma$ and any $v\in V(T_{\Gamma})$, we have $\gamma \psi_z\circ \varphi_z(v)=\gamma v=\psi_z\circ \varphi_z(\gamma v)$.
The equality $\gamma \tilde{\psi}_z\circ \tilde{\varphi}_z(s)=\tilde{\psi}_z\circ \tilde{\varphi}_z(\gamma s)$ thus holds for any $\gamma \in \Gamma$ and any $s\in T_{\Gamma}$.
We have the inclusions $\Gamma_e< \Gamma_d$ and $\Gamma_{e'}<\Gamma_d$, and thus $\Gamma_e=\Gamma_d=\Gamma_{e'}$ thanks to the assumption that $q$ is not a multiple of $p$.
There exist $k, k'\in \mathbb{Z}$ with $d=a_u^ke=a_u^{k'}e'$, where $a_u$ is a generator of $\Gamma_u$.
The equality $a_u^{k-k'}\tilde{\psi}_z\circ \tilde{\varphi}_z(e)=\tilde{\psi}_z\circ \tilde{\varphi}_z(e')$ implies the equality $a_u^{k-k'}d=d$.
We have $a_u^{k-k'}\in \Gamma_d=\Gamma_e$, and thus $e'=a_u^{k-k'}e=e$.
\end{proof}
\begin{claim}
For a.e.\ $z\in Z$, the simplicial isomorphism $\varphi_z\colon T_{\Gamma}\to T_{\Lambda}$ is orientation-preserving.
\end{claim}
\begin{proof}
Pick $u\in V(T_{\Gamma})$.
Let $e_1,\ldots, e_{|q|}\in E(T_{\Gamma})$ denote the edges whose boundaries contain $u$ as their origins.
Let $u_1,\ldots, u_{|q|}\in V(T_{\Gamma})$ be the vertices of $e_1,\ldots, e_{|q|}$ other than $u$, respectively.
Let $Z_1$ be a Borel subset of $Z$ with positive measure where the maps assigning $\varphi_z(u)$ and $\varphi_z(u_i)$ to $z$ are constant for any $i$.
We denote by $v$ and $v_i$ the values of the maps on $Z_1$, respectively, and denote by $f_i$ the edge in $E(T_{\Lambda})$ connecting $v$ with $v_i$.
The equality $\Gamma_{e_1}=\cdots =\Gamma_{e_{|q|}}$ implies the equality $(\mathcal{G}_{e_1})_{Z_1}=\cdots =(\mathcal{G}_{e_{|q|}})_{Z_1}$.
It follows that there exists a Borel subset $Z_2$ of $Z_1$ with positive measure such that $(\mathcal{H}_{f_1})_{Z_2}=\cdots =(\mathcal{H}_{f_{|q|}})_{Z_2}$.
By Lemma \ref{lem-edge-stab}, we have $\Lambda_{f_1}=\cdots =\Lambda_{f_{|q|}}$.
The edges $f_1,\ldots, f_{|q|}$ therefore have the same origin.
\end{proof}
The rest of the proof of Theorem \ref{thm-aut-t} is essentially the same as that of \cite[Theorem 4.4]{kida-ama}.
We define a Borel map $\varphi \colon Z\to {\rm Isom}(T_{\Gamma}, T_{\Lambda})$ by $\varphi(z)=\varphi_z$ for $z\in Z$.
Following the proof of \cite[Lemma 4.7]{kida-ama}, we can show the equality
\[\jmath \circ \alpha(\gamma, x)\varphi(x)=\varphi(\gamma \cdot x)\imath(\gamma)\]
for any $\gamma \in \Gamma$ and a.e.\ $x\in Z$ with $\gamma \cdot x\in Z$.
The coupling $\Sigma$ is identified with $X\times \Lambda$ as a measure space.
The group $\Gamma \times \Lambda$ acts on $X\times \Lambda$ by
\[(\gamma, \lambda)(x, \lambda')=(\gamma \cdot x, \alpha(\gamma, x)\lambda'\lambda^{-1})\]
for $\gamma \in \Gamma$, $\lambda, \lambda'\in \Lambda$ and $x\in X$.
We define a Borel map $\Phi \colon \Sigma \to {\rm Isom}(T_{\Lambda}, T_{\Gamma})$ by putting $\Phi((\gamma, \lambda)(x, e_{\Lambda}))=\imath(\gamma)\varphi(x)^{-1}\jmath(\lambda)^{-1}$ for $\gamma \in \Gamma$, $\lambda \in \Lambda$ and $x\in Z$, where $e_{\Lambda}$ denotes the neutral element of $\Lambda$.
This map $\Phi$ is shown to be well-defined and be a desired one.
We refer to the proof of \cite[Theorem 4.4]{kida-ama} for a precise argument.
Uniqueness of $\Phi$ follows from Lemma \ref{lem-dirac}.
\end{proof}
Through the translation between ME and WOE, we obtain the following:
\begin{cor}\label{cor-woe}
In Notation \ref{not-bs}, let $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ be ergodic f.f.m.p.\ actions such that
\begin{itemize}
\item the action of $\langle a^{d_0}\rangle$ on almost every ergodic component for the action $\langle a\rangle \c (X, \mu)$ is $(p_0, q_0)$-ergodic; and
\item the action of $\langle b^{c_0}\rangle$ on almost every ergodic component for the action $\langle b\rangle \c (Y, \nu)$ is $(r_0, s_0)$-ergodic.
\end{itemize}
If the actions $\Gamma \c (X, \mu)$ and $\Lambda \c (Y, \nu)$ are WOE, then $|p|=|r|$ and $|q|=|s|$.
\end{cor}
\subsection{A profinite completion of integers}
This subsection is preliminary to the next subsection.
We fix two integers $p$, $q$ with $2\leq p<q$.
Define $d_0$ as the greatest common divisor of $p$ and $q$, and set $p_0=p/d_0$ and $q_0=q/d_0$.
We assume that $q$ is not a multiple of $p$.
We thus have $1<p_0<q_0$.
Let $E$ denote the infinite cyclic group $\mathbb{Z}$.
For two non-negative integers $k$, $l$, we denote by $E_{k, l}$ the subgroup of $E$ with $[E: E_{k, l}]=d_0p_0^kq_0^l$.
We define $E_{\infty}$ as the projective limit
\[E_{\infty}=\varprojlim E/E_{k, l},\]
which is a compact unital ring.
The group $E$ is naturally identified with a dense subgroup of $E_{\infty}$.
For two non-negative integers $k$, $l$, we denote by $\bar{E}_{k, l}$ the closure of $E_{k, l}$ in $E_{\infty}$, which is equal to the kernel of the canonical homomorphism from $E_{\infty}$ onto $E/E_{k, l}$.
\begin{lem}\label{lem-limit}
In the above notation, the following assertions hold:
\begin{enumerate}
\item The additive group $\bar{E}_{0, 0}$ is torsion-free.
\item For any two non-negative integers $k$, $l$, the map $\sigma_{k, l}\colon \bar{E}_{0, 0}\to \bar{E}_{k, l}$ defined by $\sigma_{k, l}(x)=p_0^kq_0^lx$ for $x\in \bar{E}_{0, 0}$ is an isomorphism of additive groups.
\item For any continuous automorphism $\alpha$ of the additive group $E_{\infty}$, there exists a unit $r$ of $E_{\infty}$ with $\alpha(x)=rx$ for any $x\in E_{\infty}$.
\item Let $r$ be a unit of $E_{\infty}$.
If there exist non-negative integers $k$, $l$ with $rx=x$ for any $x\in \bar{E}_{k, l}$, then $d_0(r-1)=0$.
\item Conversely, if $r$ is a unit of $E_{\infty}$ with $d_0(r-1)=0$, then we have $rx=x$ for any $x\in \bar{E}_{0, 0}$.
\end{enumerate}
\end{lem}
\begin{proof}
To prove assertion (i), we pick an element $x$ of $\bar{E}_{0, 0}$ and a positive integer $m$ with $mx=0$.
Choose a sequence $\{ n_i\}_i$ in $E_{0, 0}$ converging to $x$.
The sequence $\{ mn_i\}_i$ then converges to $0$.
Let $k$ be a positive integer.
For any sufficiently large $i$, the integer $mn_i$ is divided by $d_0(p_0q_0)^k$.
It follows that for any sufficiently large $i$, the integer $n_i$ is also divided by $d_0(p_0q_0)^k$.
The sequence $\{ n_i\}_i$ therefore converges to $0$, and we obtain $x=0$.
Assertion (i) is proved.
Assertion (ii) follows from assertion (i).
In assertion (iii), putting $r=\alpha(1)$, we have the equality $\alpha(n)=nr$ for any $n\in \mathbb{Z}$.
The equality $\alpha(x)=rx$ for any $x\in E_{\infty}$ follows from the continuity of $\alpha$.
Let $r$ be a unit of $E_{\infty}$ satisfying the assumption in assertion (iv).
We then have $d_0p_0^kq_0^l(r-1)=0$ because $d_0p_0^kq_0^l$ belongs to $\bar{E}_{k, l}$.
On the other hand, $d_0(r-1)$ belongs to $\bar{E}_{0, 0}$.
We have $d_0(r-1)=0$ by assertion (i).
Assertion (iv) is proved.
In assertion (v), pick an element $x$ of $\bar{E}_{0, 0}$, which is approximated by a sequence $\{ d_0n_i\}_i$ of elements of $E_{0, 0}$ with $n_i\in E$ for any $i$.
Let $y\in E_{\infty}$ be an accumulation point of the sequence $\{ n_i\}_i$.
We then have $d_0y=x$ and $(r-1)x=(r-1)d_0y=0$.
Assertion (v) is proved.
\end{proof}
\subsection{Further reduction}
Under an assumption stronger than that in Theorem \ref{thm-aut-t}, we obtain the following stronger conclusion.
\begin{thm}\label{thm-mer}
In Notation \ref{not-bs}, let $(\Sigma, m)$ be a $(\Gamma, \Lambda)$-coupling.
We assume the following two conditions:
\begin{enumerate}
\item[(I)] The action of $\langle a^{d_0}\rangle$ on $\Sigma/\Lambda$ is ergodic.
\item[(II)] The action of $\langle b^{c_0}\rangle$ on $\Sigma/\Gamma$ is ergodic.
\end{enumerate}
Then there exists an integer $\varepsilon \in \{ \pm 1\}$ with $(p, q)=(\varepsilon r, \varepsilon s)$, that is, $\Gamma$ and $\Lambda$ are isomorphic.
\end{thm}
Theorem \ref{thm-woe} follows from this theorem.
Before the proof of Theorem \ref{thm-mer}, let us make a few comments.
By Lemma \ref{lem-erg-comp}, conditions (I) and (II) imply conditions (1) and (2) in Theorem \ref{thm-aut-t}, respectively.
It follows that there exists an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to ({\rm Isom}(T_{\Lambda}, T_{\Gamma}), \imath, \jmath)$ and that the equalities $|p|=|r|$ and $|q|=|s|$ hold.
In particular, the equality $c_0=d_0$ holds.
\begin{proof}[Proof of Theorem \ref{thm-mer}]
Because of the above argument, to prove the theorem, it is enough to deduce a contradiction under the assumption that $p>0$, $q>0$, $r=p$ and $s=-q$.
We then have $r_0=p_0$ and $s_0=-q_0$.
We fix the notation as follows.
Fix $u_0\in V(T_{\Gamma})$ and $v_0\in V(T_{\Lambda})$.
Let $E$ denote the stabilizer of $u_0$ in $\Gamma$, and let $F$ denote the stabilizer of $v_0$ in $\Lambda$.
For two non-negative integers $k$, $l$, we denote by $E_{k, l}$ the subgroup of $E$ with $[E: E_{k, l}]=d_0p_0^kq_0^l$.
Similarly, we denote by $F_{k, l}$ the subgroup of $F$ with $[F: F_{k, l}]=d_0p_0^kq_0^l$.
Define the projective limits
\[E_{\infty}=\varprojlim E_{k, l},\quad F_{\infty}=\varprojlim F_{k, l}.\]
We define $\mathfrak{I}$ as the space of continuous isomorphisms from $E_{\infty}$ onto $F_{\infty}$ as additive groups, and equip it with the standard Borel structure associated with the compact-open topology.
By Theorem \ref{thm-aut-t}, there exists an almost $(\Gamma \times \Lambda)$-equivariant Borel map $\Phi \colon \Sigma \to ({\rm Isom}(T_{\Lambda}, T_{\Gamma}), \imath, \jmath)$.
Let $\Sigma_0$ be the inverse image under $\Phi$ of the Borel subset $\{\, \varphi \in {\rm Isom}(T_{\Lambda}, T_{\Gamma})\mid \varphi(v_0)=u_0\,\}$.
The subset $\Sigma_0$ has positive measure because we have the equality $\Gamma \Sigma_0=\Sigma$ up to null sets.
As in \cite[Lemma 5.2]{kida-ama}, it is shown that $\Sigma_0$ is an $(E, F)$-coupling.
Let $X, Y\subset \Sigma_0$ be fundamental domains for the actions $F\c \Sigma_0$ and $E\c \Sigma_0$, respectively, such that $Z=X\cap Y$ has positive measure.
They are also fundamental domains for the actions $\Lambda \c \Sigma$ and $\Gamma \c \Sigma$, respectively.
We have the natural actions $\Gamma \c X$ and $\Lambda \c Y$, and set $\mathcal{G}=\Gamma \ltimes X$ and $\mathcal{H}=\Lambda \ltimes Y$.
Let $\alpha \colon \Gamma \times X\to \Lambda$ be the ME cocycle associated with $X$.
The Borel map $f\colon (\mathcal{G})_Z\to (\mathcal{H})_Z$ defined by $f(\gamma, x)=(\alpha(\gamma, x), x)$ for $(\gamma, x)\in (\mathcal{G})_Z$ is then an isomorphism.
We set $\mathcal{E}=E\ltimes X$ and $\mathcal{F}=F\ltimes Y$.
The equality $f((\mathcal{E})_Z)=(\mathcal{F})_Z$ holds by the construction of $X$ and $Y$.
For two non-negative integers $k$, $l$, we set
\[\mathcal{E}_{k, l}=E_{k, l}\ltimes X,\quad \mathcal{F}_{k, l}=F_{k, l}\ltimes Y.\]
\begin{claim}\label{claim-part}
For any two non-negative integers $k$, $l$ with $(k, l)\neq (0, 0)$, there exists a countable Borel partition $Z=\bigsqcup_nZ_n^{k, l}$ with the equality $f((\mathcal{E}_{k, l})_{Z_n^{k, l}})=(\mathcal{F}_{k, l})_{Z_n^{k, l}}$ for any $n$.
\end{claim}
\begin{proof}
Let $k$ and $l$ be non-negative integers with $(k, l)\neq (0, 0)$.
There exists a vertex $u\in V(T_{\Gamma})$ with $E\cap \Gamma_u=E_{k, l}$, where $\Gamma_u$ is the stabilizer of $u$ in $\Gamma$.
By Lemma \ref{lem-pre}, there exist a countable Borel partition $Z=\bigsqcup_n Z_n$ and $v_n\in V(T_{\Lambda})$ such that $f((\mathcal{G}_u)_{Z_n})=(\mathcal{H}_{v_n})_{Z_n}$ for any $n$, where we use the same notation as in the lemma.
It follows that $f$ sends $(\mathcal{E}_{k, l})_{Z_n}$ onto $(\mathcal{F}\cap \mathcal{H}_{v_n})_{Z_n}$ for any $n$.
Condition (I) in Theorem \ref{thm-mer} and Lemma \ref{lem-erg-comp} imply that $\mathcal{E}_{k, l}$ is ergodic.
By Lemma \ref{lem-index} (ii) and (iii), the index of $(\mathcal{E}_{k, l})_{Z_n}$ in $(\mathcal{E})_{Z_n}$ at a.e.\ $x\in Z_n$ is equal to $d_0p_0^kq_0^l$.
It follows that the index of $(\mathcal{F}\cap \mathcal{H}_{v_n})_{Z_n}$ in $(\mathcal{F})_{Z_n}$ at a.e.\ $x\in Z_n$ is also equal to the same number.
There exist non-negative integers $k'$, $l'$ with $(k', l')\neq (0, 0)$ and $\mathcal{F}\cap \mathcal{H}_{v_n}=\mathcal{F}_{k', l'}$.
Condition (II) in Theorem \ref{thm-mer} and Lemma \ref{lem-erg-comp} imply that $\mathcal{F}_{k', l'}$ is ergodic.
By Lemma \ref{lem-index} (ii) and (iii), we have $k'=k$ and $l'=l$.
The claim is proved.
\end{proof}
\noindent {\bf Construction of $\psi_0\colon Z\to \mathfrak{I}$.}
Let $k$ and $l$ be non-negative integers with $(k, l)\neq (0, 0)$.
We have a countable Borel partition $Z=\bigsqcup_n Z_n^{k, l}$ satisfying the conclusion in Claim \ref{claim-part}.
For each $n$, we have the isomorphism
\[E/E_{k, l}\to \mathcal{E}/\mathcal{E}_{k, l}\to (\mathcal{E})_{Z_n^{k, l}}/(\mathcal{E}_{k, l})_{Z_n^{k, l}}\stackrel{f}{\to}(\mathcal{F})_{Z_n^{k, l}}/(\mathcal{F}_{k, l})_{Z_n^{k, l}}\to \mathcal{F}/\mathcal{F}_{k, l}\to F/F_{k, l},\]
where we use Lemma \ref{lem-qu} in the isomorphisms other than the third one.
We define a Borel map $\psi_{k, l}\colon Z\times E/E_{k, l}\to F/F_{k, l}$ so that for any $x\in Z_n^{k, l}$, $\psi_{k, l}(x, \cdot)$ is equal to the above isomorphism.
This map is defined independently of the choice of the countable Borel partition of $Z$.
For any integers $k'$ and $l'$ with $k'\geq k$ and $l'\geq l$ and for a.e.\ $x\in Z$, we have the following commutative diagram, where the vertical arrows denote the canonical homomorphisms:
\[\xymatrix{
E/E_{k', l'} \ar[rr]^{\psi_{k', l'}(x, \cdot)} \ar[d] & & F/F_{k', l'} \ar[d]\\
E/E_{k, l} \ar[rr]^{\psi_{k, l}(x, \cdot)} & & F/F_{k, l}
}\]
We therefore obtain a Borel map $\psi_0\colon Z\to \mathfrak{I}$ such that for a.e.\ $x\in Z$ and for any non-negative integers $k$, $l$ with $(k, l)\neq (0, 0)$, the isomorphism $\psi_0(x)$ from $E_{\infty}$ onto $F_{\infty}$ induces the isomorphism $\psi_{k, l}(x, \cdot)$ from $E/E_{k, l}$ onto $F/F_{k, l}$.
\begin{claim}\label{claim-const}
The map $\psi_0\colon Z\to \mathfrak{I}$ is essentially constant.
\end{claim}
\begin{proof}
We fix $\gamma \in E$ and a Borel subset $A$ of $Z$ with positive measure such that the inclusion $\gamma \cdot A\subset Z$ holds, and $\alpha(\gamma, \cdot)$ is constant on $A$ with the value $\lambda \in F$.
Define an automorphism $U_{\gamma}$ of $\mathcal{G}$ by $U_{\gamma}(g)=(\gamma, r(g))g(\gamma, s(g))^{-1}$ for $g\in \mathcal{G}$.
Similarly, define an automorphism $V_{\lambda}$ of $\mathcal{H}$ by $V_{\lambda}(h)=(\lambda, r(h))h(\lambda, s(h))^{-1}$ for $h\in \mathcal{H}$.
The automorphism $U_{\gamma}$ preserves $\mathcal{E}$ and $\mathcal{E}_{k, l}$ for any non-negative integers $k$, $l$ with $(k, l)\neq (0, 0)$.
It follows that $U_{\gamma}$ induces an isomorphism
\[E/E_{k, l}\to (\mathcal{E})_A/(\mathcal{E}_{k, l})_A\to (\mathcal{E})_{\gamma \cdot A}/(\mathcal{E}_{k, l})_{\gamma \cdot A}\to E/E_{k, l},\]
which is the identity because $E$ is commutative.
Similarly, the automorphism $V_{\lambda}$ induces an isomorphism
\[F/F_{k, l}\to (\mathcal{F})_A/(\mathcal{F}_{k, l})_A\to (\mathcal{F})_{\lambda \cdot A}/(\mathcal{F}_{k, l})_{\lambda \cdot A}\to F/F_{k, l},\]
which is the identity.
Note that we have $\gamma \cdot A=\lambda \cdot A$ because $f$ induces the identity on $Z$.
The equality $f\circ U_{\gamma}=V_{\lambda}\circ f$ holds on $(\mathcal{G})_A$ by the definition of $f$.
We thus have $\psi_0(\gamma \cdot x)=\psi_0(x)$ for a.e.\ $x\in A$.
This equality holds for a.e.\ $x\in Z$ by the choice of $A$.
The claim follows because $\mathcal{E}$ is ergodic.
\end{proof}
\noindent {\bf Construction of $C$.}
The groups $\Gamma$ and $\Lambda$ do not act on $\mathfrak{I}$ naturally.
Let us now introduce a standard Borel space $C$ on which $\Gamma \times \Lambda$ acts and which contains a quotient of $\mathfrak{I}$.
We define $\mathfrak{C}_0$ as the set of continuous isomorphisms from a closed finite index subgroup of $E_{\infty}$ onto a closed finite index subgroup of $F_{\infty}$ as additive groups.
We say that two elements of $\mathfrak{C}_0$ are equivalent if there exists a finite index subgroup of $E_{\infty}$ on which they are equal.
Define $\mathfrak{C}$ as the set of equivalence classes in $\mathfrak{C}_0$.
For $\varphi \in \mathfrak{C}_0$, let $[\varphi]\in \mathfrak{C}$ denote the equivalence class of $\varphi$.
Lemma \ref{lem-limit} (ii) implies that we have the isomorphisms $\sigma_{1, 0}\colon \bar{E}_{0, 0}\to \bar{E}_{1, 0}$ and $\sigma_{0, 1}\colon \bar{E}_{0, 0}\to \bar{E}_{0, 1}$ defined as the multiplications by $p_0$ and $q_0$, respectively.
We define $\sigma \colon \bar{E}_{1, 0}\to \bar{E}_{0, 1}$ as the composition $\sigma_{0, 1}\circ (\sigma_{1, 0})^{-1}$.
In the presentation $\Gamma =\langle\, a, t\mid ta^pt^{-1}=a^q\,\rangle$, let $\Gamma$ act on $\mathfrak{C}$ from right as follows.
Let $a$ act on $\mathfrak{C}$ by the identity and $t$ act on $\mathfrak{C}$ by $[\varphi]t=[\varphi \circ \sigma]$ for $\varphi \in \mathfrak{C}_0$, where the composition $\varphi \circ \sigma$ is defined on a closed finite index subgroup of $\bar{E}_{1, 0}$.
Similarly, we have the isomorphisms $\tau_{1, 0}\colon \bar{F}_{0, 0}\to \bar{F}_{1, 0}$ and $\tau_{0, 1}\colon \bar{F}_{0, 0}\to \bar{F}_{0, 1}$ defined as the multiplications by $p_0$ and $q_0$, respectively.
We define $\tau \colon \bar{F}_{1, 0}\to \bar{F}_{0, 1}$ as the composition $\tau_{0, 1}\circ (\tau_{1, 0})^{-1}$.
We also have the automorphism $I$ of $F_{\infty}$ defined by $I(x)=-x$ for $x\in F_{\infty}$.
In the presentation $\Lambda =\langle\, b, u\mid ub^pu^{-1}=b^{-q}\,\rangle$, let $\Lambda$ act on $\mathfrak{C}$ from left as follows.
Let $b$ act on $\mathfrak{C}$ by the identity and $u$ act on $\mathfrak{C}$ by $u[\varphi]=[\tau \circ I \circ \varphi]$ for $\varphi \in \mathfrak{C}_0$.
This action of $\Lambda$ on $\mathfrak{C}$ commutes the action of $\Gamma$.
Let $U$ denote the group of units of $E_{\infty}$, which is a closed subset of $E_{\infty}$.
For $r\in U$, let $m_r$ denote the automorphism of the additive group $E_{\infty}$ defined by $m_r(x)=rx$ for $x\in E_{\infty}$.
The group $U$ acts on $\mathfrak{I}$ by the formula $r\varphi =\varphi \circ m_{r^{-1}}$ for $r\in U$ and $\varphi \in \mathfrak{I}$.
We define $U_0$ as the subgroup of $U$ consisting of all $r\in U$ with $d_0(r-1)=0$, which is closed in $U$.
The natural map from $\mathfrak{I}$ into $\mathfrak{C}$ induces an injective map from the quotient $\mathfrak{I}/U_0$ into $\mathfrak{C}$ by Lemma \ref{lem-limit} (iii)--(v).
We define $C_0$ as the image of this map and equip it with the standard Borel structure induced by the map.
For $n\in \mathbb{Z}$, we set $C_0\sigma^n=\{\, [\varphi \circ \sigma^n]\in \mathfrak{C}\mid [\varphi] \in C_0\,\}$.
Similarly, we define the subset $\tau^nC_0$ of $\mathfrak{C}$.
For any $\varphi \in \mathfrak{C}_0$ and any $n\in \mathbb{Z}$, the equality $[\varphi \circ \sigma^n]=[\tau^n\circ \varphi]$ holds.
It follows that for any $n\in \mathbb{Z}$, the equality $C_0\sigma^n=\tau^nC_0$ holds.
Since $[I\circ \varphi]$ belongs to $C_0$ for any $[\varphi]\in C_0$, we have
\[C_0\Gamma =\bigcup_{n\in \mathbb{Z}}C_0\sigma^n=\bigcup_{n\in \mathbb{Z}}\tau^nC_0=\Lambda C_0.\]
Let $C$ denote this subset of $\mathfrak{C}$, on which $\Gamma \times \Lambda$ acts.
We claim that the sets $C_0\sigma^n$ through $n\in \mathbb{Z}$ are mutually disjoint.
Let $\xi$ and $\eta$ denote the Haar measures on $E_{\infty}$ and $F_{\infty}$, respectively, with the total measure 1.
For any $[\varphi]\in C_0$, we have $\varphi_*\xi=\eta$ on a closed finite index subgroup of $F_{\infty}$.
On the other hand, we have $(\sigma^n)_*\xi=(q_0/p_0)^n\xi$ on a closed finite index subgroup of $E_{\infty}$.
The claim follows.
We equip $C$ with the standard Borel structure so that the actions of $\Gamma$ and $\Lambda$ on it are Borel.
\medskip
We define a Borel map $\psi \colon Z\to C_0$ as the composition of the essentially constant map $\psi_0\colon Z\to \mathfrak{I}$ with the natural map from $\mathfrak{I}$ into $\mathfrak{C}$.
Recall that we have the cocycle $\alpha \colon \Gamma \times X\to \Lambda$ associated with $X$.
We use a dot for the action of $\Gamma$ on $X$, that is, we set $\gamma \cdot x=(\gamma, \alpha(\gamma, x))x$ for $\gamma \in \Gamma$ and $x\in X$.
\begin{claim}\label{claim-psi-eq}
For any $\gamma \in \Gamma$ and a.e.\ $x\in Z$ with $\gamma \cdot x\in Z$, we have the equality $\alpha(\gamma, x)\psi(x)=\psi(\gamma \cdot x)\gamma$.
\end{claim}
\begin{proof}
We fix $\gamma \in \Gamma$ and a Borel subset $A$ of $Z$ with positive measure such that the inclusion $\gamma \cdot A\subset Z$ holds, and $\alpha(\gamma, \cdot)$ is constant on $A$ with the value $\lambda \in \Lambda$.
Define the automorphism $U_{\gamma}$ of $\mathcal{G}$ and the automorphism $V_{\lambda}$ of $\mathcal{H}$ by the same formulas as in the proof of Claim \ref{claim-const}.
The equality $f\circ U_{\gamma}=V_{\lambda}\circ f$ holds on $(\mathcal{G})_A$ by the definition of $f$.
Choose integers $n$, $K$ and $L$ such that $K>|n|$, $L>|n|$ and $\gamma E_{k, l}\gamma^{-1}=E_{k-n, l+n}$ for any $k, l\in \mathbb{Z}$ with $k\geq K$ and $l\geq L$.
For any such $k, l\in \mathbb{Z}$, the automorphism $U_{\gamma}$ induces an isomorphism
\begin{align*}
E_{K, L}/E_{k, l}\to &(\mathcal{E}_{K, L})_A/(\mathcal{E}_{k, l})_A\\
&\to (\mathcal{E}_{K-n, L+n})_{\gamma \cdot A}/(\mathcal{E}_{k-n, l+n})_{\gamma \cdot A}\to E_{K-n, L+n}/E_{k-n, l+n}.
\end{align*}
It moreover induces an isomorphism from $\bar{E}_{K, L}$ onto $\bar{E}_{K-n, L+n}$, which is equal to the restriction of $\sigma^n$.
For a.e.\ $x\in A$, the isomorphism $f\circ U_{\gamma}$ from $(\mathcal{G})_A$ onto $(\mathcal{H})_{\lambda \cdot A}$ therefore induces the element $[\psi_0(\gamma \cdot x)\circ \sigma^n]$ of $C$, which is equal to $\psi(\gamma \cdot x)\gamma$.
Similarly, the automorphism $V_{\lambda}$ induces an element of $\mathfrak{C}_0$, which is a restriction of $(\tau\circ I)^m$ for some $m\in \mathbb{Z}$.
For a.e.\ $x\in A$, the isomorphism $V_{\lambda}\circ f$ induces the element $[(\tau \circ I)^m\circ \psi_0(x)]$ of $C$, which is equal to $\lambda \psi_0(x)$.
\end{proof}
We define a Borel map $\Psi \colon \Sigma \to C$ by $\Psi((\gamma, \lambda)x)=\lambda \psi(x)\gamma^{-1}$ for $\gamma \in \Gamma$, $\lambda \in \Lambda$ and $x\in Z$.
As noted in the end of the proof of Theorem \ref{thm-aut-t}, the equality in Claim \ref{claim-psi-eq} is used to show that $\Psi$ is well-defined and is almost $(\Gamma \times \Lambda)$-equivariant.
By condition (I) in Theorem \ref{thm-mer}, the action $E\c \Sigma /\Lambda$ is ergodic.
Since the action of $E$ on $\Lambda \backslash C$ is trivial, the map from $\Sigma /\Lambda$ into $\Lambda \backslash C$ induced by $\Psi$ is essentially constant.
It follows that the image of $\Psi$ is essentially contained in $\Lambda [\alpha_0]$, where $\alpha_0$ is an element of $\mathfrak{C}_0$ such that $[\alpha_0]$ is the essential value of the map $\psi \colon Z\to C_0$.
A similar argument shows that the image of $\Psi$ is essentially contained in $[\alpha_0]\Gamma$.
We thus have the equality $\Lambda [\alpha_0]=[\alpha_0]\Gamma$.
It implies that there exists $n\in \mathbb{Z}$ with $[\tau \circ I\circ \alpha_0]=[\alpha_0\circ \sigma^n]$.
Since we have $[\alpha_0\circ \sigma^n]=[\tau^n\circ \alpha_0]$, there exists a closed finite index subgroup of $F_{\infty}$ on which $\tau^{1-n}\circ I$ is the identity.
One can find a non-zero element $x$ of $\bar{F}_{0, 0}$ such that $\tau^{1-n}\circ I(x)$ is defined and equal to $x$.
If $1-n\geq 0$, then there exists an element $y$ of $\bar{F}_{0, 0}$ with $x=p_0^{1-n}y=-q_0^{1-n}y$.
We thus have $(p_0^{1-n}+q_0^{1-n})y=0$ and obtain a contradiction by Lemma \ref{lem-limit} (i).
If $1-n<0$, then there exists an element $z$ of $\bar{F}_{0, 0}$ with $x=q_0^{n-1}z=-p_0^{n-1}z$.
We also obtain a contradiction.
\end{proof}
\begin{rem}
In Theorem \ref{thm-mer}, the conclusion is not necessarily true if we assume conditions (1) and (2) in Theorem \ref{thm-aut-t} in place of conditions (I) and (II).
A simple counterexample comes from commensurability between ${\rm BS}(p, q)$ and ${\rm BS}(p, -q)$ with $2\leq |p|\leq |q|$.
If $p$ and $q$ are integers with $2\leq |p|\leq |q|$, then for each $\varepsilon \in \{ \pm 1\}$, the subgroup $\Gamma_{\varepsilon}$ of ${\rm BS}(p, \varepsilon q)=\langle\, a, t\mid ta^pt^{-1}=a^{\varepsilon q}\, \rangle$ generated by $a$, $tat^{-1}$ and $t^2$ is of index 2, and $\Gamma_1$ and $\Gamma_{-1}$ are isomorphic.
Let $F\colon \Gamma_{-1}\to \Gamma_1$ be this isomorphism.
Let $\Gamma_1\times \Gamma_{-1}$ act on $\Gamma_1$ by $(\gamma, \delta)g=\gamma gF(\delta)^{-1}$ for $\gamma, g\in \Gamma_1$ and $\delta \in \Gamma_{-1}$.
Put $\Gamma ={\rm BS}(p, q)$ and $\Lambda ={\rm BS}(p, -q)$.
We define $\Sigma$ as the $(\Gamma, \Lambda)$-coupling defined by the action of $\Gamma \times \Lambda$ induced from the action of $\Gamma_1\times \Gamma_{-1}$ on $\Gamma_1$.
This coupling fulfills conditions (1) and (2) in Theorem \ref{thm-aut-t}, while it does not satisfy the conclusion of Theorem \ref{thm-mer} because $\Gamma$ and $\Lambda$ are not isomorphic.
\end{rem}
\subsection{A coupling with an unknown group}
The argument of this subsection heavily relies on Monod-Shalom \cite{ms}.
We say that a measure-preserving action of a discrete group $G$ on a measure space $(X, \mu)$ is {\it mildly mixing} if for any measurable subset $A$ of $X$ and any sequence $\{ g_i\}_i$ of $G$ tending to infinity with $\{ \mu(g_iA\triangle A)\}_i$ tending to $0$, we have either $\mu(A)=0$ or $\mu(X\setminus A)=0$.
\begin{thm}\label{thm-mm}
We set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|<|q|$ and denote by $d_0>0$ the greatest common divisor of $p$ and $q$.
Assume that $q$ is not a multiple of $p$.
Let $\Delta$ be a discrete group.
Let $(\Sigma, m)$ be a $(\Gamma, \Delta)$-coupling such that the action $\langle a^{d_0}\rangle \c \Sigma /\Delta$ is ergodic; and the action $\Delta \c \Sigma /\Gamma$ is mildly mixing.
We define $T$ as the Bass-Serre tree associated with $\Gamma$.
Then there exists an orientation-preserving simplicial action of $\Delta$ on $T$ such that for any $v\in V(T)$, the stabilizer of $v$ in $\Delta$ is infinite and amenable; and the actions of $\Delta$ on $V(T)$ and on $E(T)$ are both transitive.
\end{thm}
The conclusion of this theorem implies that $\Delta$ is an HNN extension of an infinite amenable group $\Delta_0$ relative to an isomorphism between a subgroup of $\Delta_0$ of index $|p|$ and a subgroup of $\Delta_0$ of index $|q|$.
\begin{proof}[Proof of Theorem \ref{thm-mm}]
We define $\Omega$ as the quotient space of $\Sigma \times \Sigma$ by the diagonal action of $\Delta$, which is a $(\Gamma, \Gamma)$-coupling.
By \cite[Lemma 6.5]{ms}, both the actions $\langle a^{d_0}\rangle \c \Omega /(\{ e\} \times \Gamma)$ and $\langle a^{d_0}\rangle \c \Omega /(\Gamma \times \{ e\})$ are ergodic.
In the application of the cited lemma, we use the assumption that the action $\Delta \c \Sigma/\Gamma$ is mildly mixing.
Thanks to Lemma \ref{lem-erg-comp}, Theorem \ref{thm-aut-t} can be applied to $\Omega$.
Combining Theorem \ref{thm-furman} and Lemma \ref{lem-dirac}, we obtain a homomorphism $\rho \colon \Delta \to {\rm Aut}(T)$ and an almost $(\Gamma \times \Delta)$-equivariant Borel map $\Phi \colon \Sigma \to ({\rm Aut}(T), \imath, \rho)$.
For each simplex $s$ of $T$, let ${\rm Stab}(s)$ denote the stabilizer of $s$ in ${\rm Aut}(T)$, and put
\[\Sigma_s=\Phi^{-1}({\rm Stab}(s)),\quad \Gamma_s=\imath^{-1}(\imath(\Gamma)\cap {\rm Stab}(s)),\quad \Delta_s=\rho^{-1}(\rho(\Delta)\cap {\rm Stab}(s)).\]
As in \cite[Lemma 5.2]{kida-ama}, we can show that $\Sigma_s$ is a $(\Gamma_s, \Delta_s)$-coupling.
As in \cite[Lemma 5.3]{kida-ama}, using ergodicity of the action $\Gamma_s \c \Sigma/\Delta$ for each simplex $s$ of $T$, we can show that the actions of $\Delta$ on $V(T)$ and on $E(T)$ are both transitive.
\end{proof}
\section{Non-trivial orbit equivalence}\label{sec-oe}
This section is devoted to the proof of the following:
\begin{thm}\label{thm-woe-non-conj}
Let $p$ and $q$ be integers with $2\leq |p|\leq |q|$, and set $\Gamma ={\rm BS}(p, q)$.
Then there exist two f.f.m.p.\ actions $\Gamma \c (X, \mu)$ and $\Gamma \c (Y, \nu)$ such that they are WOE, but are not conjugate; and the actions of any non-trivial elliptic subgroup of $\Gamma$ on $(X, \mu)$ and on $(Y, \nu)$ are both ergodic.
\end{thm}
The two actions of $\Gamma$ in this theorem are not virtually conjugate either in the sense of \cite[Definition 1.3]{kida-oer} because $\Gamma$ is torsion-free by \cite[Theorem IV.2.4]{ls} and because the actions of any finite index subgroup of $\Gamma$ on $(X, \mu)$ and on $(Y, \nu)$ are ergodic.
It follows that under the assumption in Theorem \ref{thm-mer}, we cannot conclude that the two actions $\Gamma \c \Sigma /\Lambda$ and $\Lambda \c \Sigma /\Gamma$ in it are virtually conjugate.
Until Theorem \ref{thm-sigma}, we set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$.
Let $d_0>0$ denote the greatest common divisor of $p$ and $q$, and set $p_0=p/d_0$ and $q_0=q/d_0$.
We fix a presentation $\Gamma =\langle \, a, t\mid ta^pt^{-1}=a^q\,\rangle$.
Let $T$ be the Bass-Serre tree associated with $\Gamma$.
Let $\imath \colon \Gamma \to {\rm Aut}(T)$ be the homomorphism associated with the action of $\Gamma$ on $T$.
\subsection{A standard coupling}
We define a continuous homomorphism $\tau \colon {\rm Aut}(T)\to \mathbb{Z}$ as follows.
Let $E_+$ and $E_-$ denote the sets of oriented edges of $T$ introduced in Section \ref{sec-bs}.
We set $\sigma(e)=1$ for each $e\in E_+$, and set $\sigma(f)=-1$ for each $f\in E_-$.
Fix a vertex $v_0$ of $T$.
For $\varphi \in {\rm Aut}(T)$, let $e_1,\ldots, e_n$ be the shortest sequence of oriented edges of $T$ such that the origin of $e_1$ is $v_0$; the terminal of $e_n$ is $\varphi(v_0)$; and for any $i=1,\ldots, n-1$, the terminal of $e_i$ and the origin of $e_{i+1}$ are equal.
We then define $\tau(\varphi)$ as the integer $\sum_{i=1}^n\sigma(e_i)$.
The map $\tau$ is independent of the choice of $v_0$ and is indeed a continuous homomorphism.
Note that $\tau(\imath(a))=0$ and $\tau(\imath(t))=1$.
We set $L=\mathbb{R}$.
Let ${\rm Aut}(T)$ act on $L$ by $\varphi x=(q/p)^{\tau(\varphi)}x$ for $\varphi \in {\rm Aut}(T)$ and $x\in L$.
We then have the semi-direct product $L\rtimes {\rm Aut}(T)$.
Let $K$ denote the closure of the cyclic group $\langle \imath(a)\rangle$ in ${\rm Aut}(T)$, which is compact and abelian.
We define $G$ as the subgroup of $L\rtimes {\rm Aut}(T)$ generated by $L$, $K$ and $\imath(t)$.
The subgroup $G$ is closed in $L\rtimes {\rm Aut}(T)$.
For each $\theta \in L\setminus \{ 0\}$, we define a homomorphism $\pi_{\theta}\colon \Gamma \to G$ by $\pi_{\theta}(a)=(\theta, \imath(a))$ and $\pi_{\theta}(t)=(0, \imath(t))$.
This is injective, and the image $\pi_{\theta}(\Gamma)$ is a cocompact lattice in $G$ such that $[0, |\theta|)\times K$ is its fundamental domain.
Let $\alpha_{\theta}$ be the automorphism of $L\rtimes {\rm Aut}(T)$ that is the identity on ${\rm Aut}(T)$ and is the multiplication by $\theta$ on $L$.
The equality $\pi_{\theta}=\alpha_{\theta}\circ \pi_1$ on $\Gamma$ then holds.
It follows that $(G, \pi_{\theta}, \pi_1)$ is a $(\Gamma, \Gamma)$-coupling.
In the rest of this subsection, we examine the action of $\Gamma$ on $G/\pi_1(\Gamma)$ through $\pi_{\theta}$ and that on $\pi_{\theta}(\Gamma)\backslash G$ through $\pi_1$.
The argument in the previous paragraph implies that the latter action is isomorphic to the action of $\Gamma$ on $G/\pi_1(\Gamma)$ through $\pi_{\theta^{-1}}$.
Since $L$ and $K$ commute in $G$, we identify the subgroup of $G$ generated by them with the direct product $L\times K$.
We set $E=\langle a\rangle$.
The space $G/\pi_1(\Gamma)$ is then identified with $(L\times K)/\pi_1(E)$.
We denote the latter space by $M$, which is a compact abelian group.
\begin{lem}\label{lem-dense}
Let $\theta$ be an irrational number.
Then the action of any non-trivial subgroup of $E$ on $M$ through $\pi_{\theta}$ is ergodic.
\end{lem}
\begin{proof}
Let $n$ be a non-zero integer.
Let $\Lambda$ denote the subgroup of $L\times K$ generated by $\pi_{\theta}(\langle a^n\rangle )$ and $\pi_1(E)$.
It suffices to show that $\Lambda$ is dense in $L\times K$.
The group $\Lambda$ is generated by $(n(\theta -1), e)$ and $(1, \imath(a))$.
Density of $\Lambda$ in $L\times K$ is thus equivalent to density of the subgroup generated by $([1], \imath(a))$ in the group $\mathbb{R}/n(\theta -1)\mathbb{Z}\times K$, where for $r\in \mathbb{R}$, $[r]$ denotes the equivalence class of $r$ in $\mathbb{R}/n(\theta -1)\mathbb{Z}$.
The latter condition follows from the assumption that $\theta$ is irrational.
\end{proof}
\begin{prop}\label{prop-conj}
For each $i=1, 2$, let $B_i$ be a compact, second countable and abelian group with the normalized Haar measure $m_i$, and let $\Lambda_i$ be a dense countable subgroup of $B_i$.
Let $\Lambda_i$ act on $(B_i, m_i)$ by left multiplication.
We suppose that the two actions $\Lambda_1\c (B_1, m_1)$ and $\Lambda_2\c (B_2, m_2)$ are conjugate, that is, we have an isomorphism $F\colon \Lambda_1\to \Lambda_2$ and a Borel isomorphism $f$ from a conull Borel subset of $B_1$ onto a conull Borel subset of $B_2$ with $f_*m_1=m_2$ and $f(\lambda b)=F(\lambda)f(b)$ for any $\lambda \in \Lambda_1$ and a.e.\ $b\in B_1$.
Then there exist a continuous isomorphism $\bar{F}\colon B_1\to B_2$ and an element $b_0\in B_2$ such that $\bar{F}$ extends $F$ and the equality $f(b)=\bar{F}(b)b_0$ holds for a.e.\ $b\in B_1$.
\end{prop}
\begin{proof}
We define a Borel map $h\colon B_1\times B_1\to B_2$ by $h(b, c)=f(b)f(c)f(bc)^{-1}$ for $b, c\in B_1$.
Since $B_1$ and $B_2$ are abelian, we have $h(\lambda b, c)=h(b, \lambda c)=h(b, c)$ for any $\lambda \in \Lambda_1$ and a.e.\ $(b, c)\in B_1\times B_1$.
Ergodicity of the action of $\Lambda_1\times \Lambda_1$ on $B_1\times B_1$ by left multiplication implies that $h$ is essentially constant.
Let $b_0\in B_2$ denote the essential image of $h$.
We define a Borel map $\bar{f}\colon B_1\to B_2$ by $\bar{f}(b)=f(b)b_0^{-1}$ for $b\in B_1$.
The map $\bar{f}$ then preserves products a.e.
It follows from \cite[Theorems B.2 and B.3]{zim-book} that there exists a continuous homomorphism $\bar{F}\colon B_1\to B_2$ with $\bar{F}=\bar{f}$ a.e.
For any $\lambda \in \Lambda_1$ and a.e.\ $b\in B_1$, we have
\[F(\lambda)f(b)=f(\lambda b)=\bar{F}(\lambda b)b_0=\bar{F}(\lambda)\bar{F}(b)b_0=\bar{F}(\lambda)f(b)\]
and thus $F(\lambda)=\bar{F}(\lambda)$.
The map $\bar{F}$ extends $F$ and is surjective.
An argument for $f^{-1}$ in place of $f$ implies that $\bar{F}$ is an isomorphism.
\end{proof}
Let $\eta \colon L\times K\to M$ denote the canonical projection.
\begin{lem}\label{lem-path}
For any $k\in K$, $\eta(L\times \{ k\})$ is a path-connected component of $M$.
\end{lem}
\begin{proof}
Let $x$ be an element of $L$.
We define $I$ as the closed interval $[x-1/3, x+1/3]$ in $L$.
The restriction of $\eta$ to $I\times K$ is injective and is thus a homeomorphism onto the image.
For any $k\in K$, the image $\eta(I\times K)$ contains an open neighborhood of $\eta(x, k)$ in $M$ because $\eta$ is an open map.
Since $K$ is totally disconnected, for any $k\in K$ and any continuous path $l\colon [0, 1]\to M$ with $l(t_0)=\eta(x, k)$ for some $t_0\in [0, 1]$, there exists a neighborhood of $t_0$ in $[0, 1]$ such that $l$ sends any element of it to $\eta(I\times \{ k\})$.
It follows that for any continuous map from $[0, 1]$ into $M$, there exists $k\in K$ such that the image of the map is contained in $\eta(L\times \{ k\})$.
\end{proof}
\begin{lem}\label{lem-rat}
Let $F$ be a continuous automorphism of $M$.
Let $\alpha$ be the continuous automorphism of $L$ with $F(\eta(x, e))=\eta(\alpha(x), e)$ for any $x\in L$.
Then the element of $L$, $\alpha(1)$, is a rational number.
\end{lem}
By Lemma \ref{lem-path}, for any continuous automorphism $F$ of $M$, there exists a unique continuous automorphism $\alpha$ of $L$ satisfying the equality $F(\eta(x, e))=\eta(\alpha(x), e)$ for any $x\in L$.
\begin{proof}[Proof of Lemma \ref{lem-rat}]
For any positive integer $n$, we have the equality $\eta(d_0p_0^nq_0^n, e)=\eta(0, \imath(a)^{-d_0p_0^nq_0^n})$.
The right hand side approaches $0$ as $n$ goes to infinity because $\imath(a)^{-d_0p_0^nq_0^n}$ approaches the identity in ${\rm Aut}(T)$.
By continuity of $F$, the element $F(\eta(d_0p_0^nq_0^n, e))=\eta(d_0p_0^nq_0^n\alpha(1), e)$ also approaches $0$.
Let $\eta_1\colon \mathbb{R}\to \mathbb{R}/\mathbb{Z}$ denote the canonical projection.
It follows that $\eta_1(d_0p_0^nq_0^n\alpha(1))$ approaches $0$ as $n$ goes to infinity.
If $|p_0|=|q_0|=1$, then $d_0\alpha(1)$ is an integer, and thus $\alpha(1)$ is rational.
Otherwise, considering the $|p_0q_0|$-adic expansion of $d_0\alpha(1)$, we find integers $r$, $n$ with $n\geq 0$ and $d_0\alpha(1)=r/|p_0q_0|^n$.
The number $\alpha(1)$ is thus rational.
\end{proof}
\begin{lem}\label{lem-conj-e}
Let $\theta$ and $\theta'$ be irrational numbers.
If the action of $E$ on $M$ through $\pi_{\theta}$ and that through $\pi_{\theta'}$ are conjugate, then $(\theta'-1)/(\theta-1)$ is a rational number.
\end{lem}
\begin{proof}
The restriction of $\pi_{\theta}$ to $E$ is injective, and $\pi_{\theta}(E)$ is dense in $M$ by Lemma \ref{lem-dense}.
The same property holds for $\theta'$.
Proposition \ref{prop-conj} implies that there exists a continuous automorphism $F$ of $M$ such that either $F\circ \pi_{\theta}=\pi_{\theta'}$ on $E$ or $F\circ \pi_{\theta}=\pi_{\theta'}\circ I$ on $E$, where $I$ is the automorphism of $E$ defined by $I(x)=x^{-1}$ for $x\in E$.
Let $\alpha$ be the continuous automorphism of $L$ with the equality $F(\eta(x, e))=\eta(\alpha(x), e)$ for any $x\in L$.
If the equality $F\circ \pi_{\theta}=\pi_{\theta'}$ on $E$ holds, then we have $F(\eta(\theta, \imath(a)))=\eta(\theta', \imath(a))$ and thus $F(\eta(\theta-1, e))=\eta(\theta'-1, e)$.
It follows that $\alpha(\theta-1)=\theta'-1$.
Lemma \ref{lem-rat} implies that $(\theta'-1)/(\theta-1)$ is rational.
A similar argument can be applied to the other case.
\end{proof}
\begin{thm}\label{thm-not-conj}
Let $\theta$ be an irrational number.
Then the following assertions hold:
\begin{enumerate}
\item The action of $E$ on $(L\times K)/\pi_1(E)$ through $\pi_{\theta}$ and the action of $E$ on $\pi_{\theta}(E)\backslash (L\times K)$ through $\pi_1$ are not conjugate.
\item The action of $\Gamma$ on $G/\pi_1(\Gamma)$ through $\pi_{\theta}$ and the action of $\Gamma$ on $\pi_{\theta}(\Gamma)\backslash G$ through $\pi_1$ are not conjugate.
\end{enumerate}
\end{thm}
\begin{proof}
As noted right before Lemma \ref{lem-dense}, the action of $E$ on $\pi_{\theta}(E)\backslash (L\times K)$ through $\pi_1$ is conjugate with the action of $E$ on $(L\times K)/\pi_1(E)$ through $\pi_{\theta^{-1}}$.
Since we have the equality $(\theta^{-1}-1)/(\theta-1)=-\theta^{-1}$, assertion (i) follows from Lemma \ref{lem-conj-e}.
Any automorphism of $\Gamma$ preserves elliptic subgroups of $\Gamma$ because ellipticity of elements of $\Gamma$ is algebraic as noticed in the third paragraph in Section \ref{sec-bs}.
Assertion (ii) therefore follows from assertion (i).
\end{proof}
Finally, we notice that the action of $\Gamma$ on $G/\pi_1(\Gamma)$ through $\pi_{\theta}$ is not essentially free.
Let $\mathfrak{m} \colon \Gamma \to \mathbb{R}_+^{\times}$ denote the modular homomorphism.
For any $\theta \in L\setminus \{ 0\}$ and $\gamma \in \Gamma$, let $l_{\theta}(\gamma)\in L$ denote the first coordinate of $\pi_{\theta}(\gamma)$ in $L\rtimes {\rm Aut}(T)$.
Since $\imath(\ker \mathfrak{m})$ acts on $L$ trivially, the map $l_{\theta}\colon \ker \mathfrak{m} \to L$ is a homomorphism.
Let $N$ denote the kernel of this homomorphism.
The group $N$ is independent of $\theta$ because we have the equality $l_{\theta}(\gamma)=\theta l_1(\gamma)$ for any $\gamma \in \Gamma$.
Define the derived subgroups
\[\Gamma^{(1)}=[\Gamma, \Gamma],\quad \Gamma^{(2)}=[\Gamma^{(1)}, \Gamma^{(1)}].\]
We then have the inclusions $\Gamma^{(1)}<\ker \mathfrak{m}$ and $\Gamma^{(2)}<N$.
\begin{lem}\label{lem-trivial}
For any $\theta \in L\setminus \{ 0\}$, the action of $N$ on $G/\pi_1(\Gamma)$ through $\pi_{\theta}$ is trivial.
\end{lem}
\begin{proof}
Pick $\gamma \in N$.
Choose a positive integer $n$ such that $\gamma$ commutes $a^n$.
Let $K_1$ denote the closure of $\langle \imath(a^n)\rangle$ in $K$, which is a subgroup of $K$ of index at most $n$.
The subset $[0, 1)\times K$ of $G$ is a fundamental domain for the action of $\Gamma$ on $G$ by right multiplication through $\pi_1$.
For any $x\in [0, 1)$ and $k\in K$, choosing $m\in \mathbb{Z}$ and $k_1\in K_1$ with $k=\imath(a^m)k_1$, we have the equality
\begin{align*}
\pi_{\theta}(\gamma)(x, k)&=(x, \imath(\gamma)k)=(x, k_1\imath(\gamma)\imath(a^m))=(x, k_1\imath(a^m)\imath(a^{-m}\gamma a^m))\\
&=(x, k\imath(a^{-m}\gamma a^m))=(x, k)\pi_1(a^{-m}\gamma a^m),
\end{align*}
where we use commutativity of $K$, and the first and last equalities hold because $\gamma$ and $a^{-m}\gamma a^m$ lie in $N$.
The lemma therefore follows.
\end{proof}
\subsection{A modified coupling}
Modifying the coupling $(G, \pi_{\theta}, \pi_1)$, we obtain a WOE between f.f.m.p.\ actions of $\Gamma$ satisfying the conclusion in Theorem \ref{thm-woe-non-conj}.
We pick an essentially free and measure-preserving action of $\Gamma$ on a standard probability space $(Z, \xi)$.
Let $(\Sigma, m)$ denote the product space $(Z\times G, \xi \times m_G)$, where $m_G$ is the Haar measure on $G$.
We define an action of $\Gamma \times \Gamma$ on $\Sigma$ by
\[(\gamma_1, \gamma_2)(z, g)=(\gamma_1z, \pi_{\theta}(\gamma_1)g\pi_1(\gamma_2)^{-1})\]
for $\gamma_1, \gamma_2\in \Gamma$, $z\in Z$ and $g\in G$.
This action makes $(\Sigma, m)$ into a $(\Gamma, \Gamma)$-coupling.
Let $\mathsf{L}, \mathsf{R}\colon \Gamma \to \Gamma \times \Gamma$ be the homomorphisms defined by
\[\mathsf{L}(\gamma)=(\gamma, e),\quad \mathsf{R}(\gamma)=(e, \gamma)\]
for $\gamma \in \Gamma$.
We define two subsets $X_0$, $Y_0$ of $G$ and two subsets $X$, $Y$ of $\Sigma$ by
\[X_0=[0, 1)\times K,\quad Y_0=[0, |\theta|)\times K,\quad X=Z\times X_0,\quad Y=Z\times Y_0.\]
The subsets $X_0$ and $Y_0$ are fundamental domains for the actions $\mathsf{R}(\Gamma)\c G$ and $\mathsf{L}(\Gamma)\c G$, respectively.
The subsets $X$ and $Y$ are fundamental domains for the actions $\mathsf{R}(\Gamma)\c \Sigma$ and $\mathsf{L}(\Gamma)\c \Sigma$, respectively.
We have the natural actions $\Gamma \c X$ and $\Gamma \c Y$.
The former action is isomorphic to the diagonal action of $\Gamma$ on $Z\times X_0$.
It follows that the actions $\Gamma \c X$ and $\Gamma \c Y$ are essentially free because so is the action $\Gamma \c Z$.
Let $\beta \colon \Gamma \times Y_0\to \Gamma$ be the ME cocycle for the coupling $(G, \pi_{\theta}, \pi_1)$.
As for the action $\Gamma \c Y$, we have the equality
\[\gamma \cdot (z, y)=(\beta(\gamma, y)z, \gamma \cdot y)\]
for any $\gamma \in \Gamma$, any $z\in Z$ and a.e.\ $y\in Y_0$.
We now prove that the actions $\Gamma \c X$ and $\Gamma \c Y$ satisfy the desired property in Theorem \ref{thm-woe-non-conj}.
\begin{lem}\label{lem-z-conj}
Suppose that $\theta$ is irrational and that any non-neutral element of $\Gamma^{(2)}$ acts on $(Z, \xi)$ ergodically.
Then the following assertions hold:
\begin{enumerate}
\item For any non-neutral element $\gamma_0$ of $\Gamma^{(2)}$, the ergodic decomposition for the action $\langle \gamma_0 \rangle \c X$ is equal to the projection from $X=Z\times X_0$ onto $X_0$.
Moreover, the ergodic decomposition for the action $\langle \gamma_0 \rangle \c Y$ is equal to the projection from $Y=Z\times Y_0$ onto $Y_0$.
\item The actions $\Gamma \c X$ and $\Gamma \c Y$ are not conjugate.
\end{enumerate}
\end{lem}
\begin{proof}
The former assertion in assertion (i) holds because $\Gamma$ acts on $X=Z\times X_0$ diagonally and because $\Gamma^{(2)}$ acts on $X_0$ trivially by Lemma \ref{lem-trivial}.
We prove the latter assertion.
Pick a non-neutral element $\gamma_0$ of $\Gamma^{(2)}$.
As in the second paragraph of the proof of Lemma \ref{lem-trivial}, for any $x\in [0, |\theta|)$ and $k\in K$, we can find $m\in \mathbb{Z}$ such that $\beta(\gamma_0^n, (x, k))=a^m\gamma_0^na^{-m}$ for any $n\in \mathbb{Z}$.
It follows that the restriction of the action $\langle \gamma_0 \rangle \c Y$ to $Z\times \{ (x, k)\}$ is isomorphic to the action $\langle \gamma_0 \rangle \c Z$, which is ergodic by assumption.
The latter assertion in assertion (i) is proved.
We prove assertion (ii).
Assuming that the two actions $\Gamma \c X$ and $\Gamma \c Y$ are conjugate, we deduce a contradiction.
Let $F$ be an automorphism of $\Gamma$ and $f$ a Borel isomorphism from a conull Borel subset of $X$ onto a conull Borel subset of $Y$ such that $f$ preserves the classes of the measures on $X$ and $Y$ and satisfies the equality $f(\gamma \cdot x)=F(\gamma)\cdot f(x)$ for any $\gamma \in \Gamma$ and a.e.\ $x\in X$.
Pick a non-neutral element $\gamma_0$ of $\Gamma^{(2)}$.
Since $F$ preserves $\Gamma^{(2)}$, the element $F(\gamma_0)$ is also non-neutral and belongs to $\Gamma^{(2)}$.
Assertion (i) implies that we have a Borel isomorphism $h$ from a conull Borel subset of $X_0$ onto a conull Borel subset of $Y_0$ which preserves the classes of the measures on $X_0$ and $Y_0$ and satisfies the equality $f(Z\times \{ x\})=Z\times \{ h(x)\}$ up to null sets, for a.e.\ $x\in X_0$.
For any $\gamma \in \Gamma$ and a.e.\ $x\in X_0$, we then have the equality $h(\gamma \cdot x)=F(\gamma)\cdot h(x)$.
It follows that the actions $\Gamma \c X_0$ and $\Gamma \c Y_0$ are conjugate.
This contradicts Theorem \ref{thm-not-conj}.
\end{proof}
\begin{lem}\label{lem-z-erg}
Suppose that $\theta$ is irrational.
Let $H$ be a non-trivial subgroup of $E$.
Then the following assertions hold:
\begin{enumerate}
\item We identify $E$ with $\mathbb{Z}$ through the isomorphism sending $a$ to $1$.
Then for a.e.\ $y\in Y_0$, the function $\beta(\cdot, y)$ from $H$ into $E$ is non-decreasing if $\theta$ is positive, and is non-increasing if $\theta$ is negative.
Moreover, for a.e.\ $y\in Y_0$, the function $\beta(\cdot, y)$ is bounded neither below nor above.
\item If the action of $E$ on $(Z, \xi)$ is mixing, then the actions of $H$ on $X$ and on $Y$ are both ergodic.
\end{enumerate}
\end{lem}
\begin{proof}
For any $x\in [0, |\theta|)$, any $k\in K$ and any $n\in \mathbb{Z}$, the integer $m$ satisfying the equation $\beta(a^n, (x, k))=a^m$ is determined by the condition $x-n+\theta m\in [0, |\theta|)$.
Assertion (i) follows.
In assertion (ii), ergodicity of the action of $H$ on $X$ follows from Lemma \ref{lem-dense} and the assumption that the action of $E$ on $(Z, \xi)$ is mixing.
We show that the action of $H$ on $Y$ is ergodic in the case where $\theta$ is positive.
The proof of the other case is similar and is thus omitted.
Let $\nu_0$ denote the measure on $Y_0$.
We may assume that $\nu_0$ is a probability measure, and define the probability measure $\nu =\xi \times \nu_0$ on $Y$.
Let $b$ be the generator of $H$ that is a positive power of $a$.
It is enough to show that for any Borel subsets $A_1, A_2\subset Z$ and $B_1, B_2\subset Y_0$, we have
\[\frac{1}{n}\sum_{k=1}^n\nu(b^k(A_1\times B_1)\cap (A_2\times B_2))\to \nu(A_1\times B_1)\nu(A_2\times B_2)\]
as $n$ goes to infinity (see \cite[I.2.6 (iii)]{bm}).
For non-negative integers $k$, $m$, we set
\[C_k^m=\{\, y\in B_1\mid \beta(b^k, y)=a^m\,\}.\]
For any positive integer $k$, we have the decomposition $B_1=\bigsqcup_{m=0}^{\infty}C_k^m$ and thus
\[b^k(A_1\times B_1)=\bigsqcup_{m=0}^{\infty}(a^mA_1)\times (b^kC_k^m).\]
Pick a positive real number $\varepsilon$.
Since the action of $E$ on $(Z, \xi)$ is mixing, there exists a positive integer $M_1$ such that for any integer $m$ with $m\geq M_1$, we have
\[|\xi(a^m A_1\cap A_2)-\xi(A_1)\xi(A_2)|<\varepsilon.\]
By assertion (i), there exists a positive integer $M_2$ such that for any integer $k$ with $k\geq M_2$, we have $\sum_{m=0}^{M_1-1}\nu_0(C_k^m)<\varepsilon$.
For any integer $k$ with $k\geq M_2$, we have
\begin{align*}
&\, \left|\,\nu(b^k(A_1\times B_1)\cap (A_2\times B_2))-\xi(A_1)\xi(A_2)\nu_0(b^kB_1\cap B_2)\,\right|\\
= & \,\left|\, \sum_{m=0}^{\infty}\xi(a^mA_1\cap A_2)\nu_0(b^kC_k^m\cap B_2)-\xi(A_1)\xi(A_2)\nu_0(b^kB_1\cap B_2)\,\right|\\
\leq & \,\varepsilon +\left|\, \sum_{m=M_1}^{\infty}\xi(a^mA_1\cap A_2)\nu_0(b^kC_k^m\cap B_2)-\xi(A_1)\xi(A_2)\nu_0(b^kB_1\cap B_2)\,\right|\\
\leq & \,2\varepsilon +\xi(A_1)\xi(A_2)\left|\, \sum_{m=M_1}^{\infty}\nu_0(b^kC_k^m\cap B_2)-\nu_0(b^kB_1\cap B_2)\,\right| \leq 3\varepsilon.
\end{align*}
For any sufficiently large integer $n$ with $n\geq M_2$ and $n\geq M_2/\varepsilon$, we have
\begin{align*}
&\, \left|\,\frac{1}{n}\sum_{k=1}^n\nu(b^k(A_1\times B_1)\cap (A_2\times B_2))-\nu(A_1\times B_1)\nu(A_2\times B_2)\,\right|\\
\leq & \, \left|\,\frac{1}{n}\sum_{k=1}^{M_2-1}\nu(b^k(A_1\times B_1)\cap (A_2\times B_2))\,\right|\\
& +\left|\,\frac{1}{n}\sum_{k=M_2}^n\nu(b^k(A_1\times B_1)\cap (A_2\times B_2))-\xi(A_1)\xi(A_2)\nu_0(B_1)\nu_0(B_2)\,\right|\\
\leq & \, \varepsilon +3\varepsilon + \xi(A_1)\xi(A_2)\left|\,\frac{1}{n}\sum_{k=M_2}^n \nu_0(b^kB_1\cap B_2)-\nu_0(B_1)\nu_0(B_2)\,\right|\\
\leq & \, 4\varepsilon + \xi(A_1)\xi(A_2)\frac{1}{n}\sum_{k=1}^{M_2-1}\nu_0(b^kB_1\cap B_2)\\
& +\xi(A_1)\xi(A_2)\left|\, \frac{1}{n}\sum_{k=1}^n\nu_0(b^kB_1\cap B_2)-\nu_0(B_1)\nu_0(B_2)\,\right|\\
\leq &\, 4\varepsilon +\varepsilon +\varepsilon =6\varepsilon,
\end{align*}
where we use ergodicity of the action of $H$ on $Y_0$ in the last inequality, which follows from Lemma \ref{lem-dense}.
\end{proof}
Combining Lemmas \ref{lem-z-conj} and \ref{lem-z-erg}, we obtain the following:
\begin{thm}\label{thm-sigma}
If $\theta$ is irrational and the action $\Gamma \c (Z, \xi)$ is mixing, then the $(\Gamma, \Gamma)$-coupling $(\Sigma, m)$ satisfies the following three properties:
\begin{itemize}
\item The action $\Gamma \times \Gamma \c (\Sigma, m)$ is essentially free.
\item The two actions $\mathsf{L}(\Gamma)\c \Sigma/\mathsf{R}(\Gamma)$ and $\mathsf{R}(\Gamma)\c \Sigma/\mathsf{L}(\Gamma)$ are not conjugate.
\item For any non-trivial elliptic subgroup $H$ of $\Gamma$, the actions $\mathsf{L}(H)\c \Sigma /\mathsf{R}(\Gamma)$ and $\mathsf{R}(H)\c \Sigma/\mathsf{L}(\Gamma)$ are both ergodic.
\end{itemize}
\end{thm}
Theorem \ref{thm-woe-non-conj} is a direct consequence of this theorem.
\begin{rem}\label{rem-ber}
In contrast with Theorem \ref{thm-woe-non-conj}, we obtain rigidity of Bernoulli shifts of Baumslag-Solitar groups as an application of Popa's cocycle superrigidity theorem \cite{popa-gap}.
Let $\Gamma$ be a discrete group.
An infinite subgroup $H$ of $\Gamma$ is called {\it wq-normal} in $\Gamma$ if for any subgroup $H_1$ with $H<H_1<\Gamma$ and $H_1\neq \Gamma$, there exists an element $\gamma \in \Gamma \setminus H_1$ with $\gamma H_1\gamma^{-1}\cap H_1$ infinite.
This definition is different from that in \cite{popa-gap}.
Equivalence between these two definitions is discussed in \cite[Definition 2.3]{popa-coh}.
\begin{lem}\label{lem-wq}
We set $\Gamma ={\rm BS}(p, q)=\langle\, a, t\mid ta^pt^{-1}=a^q\,\rangle$ with $2\leq |p|\leq |q|$.
Then $\langle a^q\rangle$ is wq-normal in $\Gamma$, and the centralizer of $\langle a^q\rangle$ in $\Gamma$ is non-amenable.
\end{lem}
\begin{proof}
Let $H_1$ be a proper subgroup of $\Gamma$ containing $\langle a^q\rangle$. Either $a$ or $t$ does not belong to $H_1$.
Both $aH_1a^{-1}\cap H_1$ and $tH_1t^{-1}\cap H_1$ are infinite because we have $a\langle a^q\rangle a^{-1}=\langle a^q\rangle$ and $\langle a^{q^2}\rangle <t\langle a^q\rangle t^{-1}\cap \langle a^q\rangle$.
The former assertion of the lemma is proved.
If $|p|=|q|$, then the centralizer of $\langle a^q\rangle$ in $\Gamma$ contains $a$, $tat^{-1}$ and $t^2$ and is of finite index in $\Gamma$.
If $|p|<|q|$, then the centralizer of $\langle a^q\rangle$ in $\Gamma$ contains $a$ and $tat^{-1}$ and is non-amenable.
The latter assertion of the lemma follows.
\end{proof}
For a discrete group $\Gamma$ and a standard probability space $(X_0, \mu_0)$ such that there is no point of $X_0$ whose measure is equal to 1, the {\it Bernoulli shift} $\Gamma \c (X_0, \mu_0)^{\Gamma}$ is defined by the formula $\gamma (x_{\delta})_{\delta \in \Gamma}=(x_{\gamma^{-1}\delta})_{\delta \in \Gamma}$ for $\gamma \in \Gamma$ and $(x_{\delta})_{\delta \in \Gamma}\in X_0^{\Gamma}$.
Lemma \ref{lem-wq} and \cite[Corollary 1.2]{popa-gap} imply the following rigidity.
\begin{thm}\label{thm-ber}
Set $\Gamma ={\rm BS}(p, q)$ with $2\leq |p|\leq |q|$.
Let $(X_0, \mu_0)$ be a standard probability space such that there is no point of $X_0$ whose measure is equal to $1$.
If the Bernoulli shift $\Gamma \c (X_0, \mu_0)^{\Gamma}$ is WOE to an ergodic f.f.m.p.\ action $\Lambda \c (Y, \nu)$ of a discrete group $\Lambda$, then there exist a finite index subgroup $\Lambda_0$ of $\Lambda$ and a $\Lambda_0$-invariant Borel subset $Y_0$ of $Y$ satisfying the following three conditions:
\begin{itemize}
\item The equality $\nu(Y_0)/\nu(Y)=[\Lambda :\Lambda_0]^{-1}$ holds.
\item The action $\Lambda \c (Y, \nu)$ is induced from the action $\Lambda_0\c (Y_0, \nu|_{Y_0})$.
\item The actions $\Gamma \c (X_0, \mu_0)^{\Gamma}$ and $\Lambda_0\c (Y_0, \nu|_{Y_0})$ are conjugate.
\end{itemize}
\end{thm}
In \cite[Corollary 1.7]{bowen}, Bowen shows that the entropy of the base space $(X_0, \mu_0)$ is a complete conjugacy-invariant of Bernoulli shifts of a sofic Ornstein group.
The group ${\rm BS}(p, q)$ is sofic (see \cite[Example 4.6]{pestov}) and is also Ornstein because it has an infinite amenable subgroup which is Ornstein (see \cite[p.218]{bowen}).
Theorem \ref{thm-ber} implies that the entropy of $(X_0, \mu_0)$ is also a complete WOE-invariant for Bernoulli shifts of ${\rm BS}(p, q)$.
\end{rem}
|
\section{Introduction}
Phase transitions in QCD have been extensively studied in lattice
quantum chromodynamics. While recent development enables us to
perform numerical simulations at physical quark masses, which revealed a
crossover nature of the QCD phase transition at finite temperature \cite{aoki06},
analyses at nonzero quark chemical potential $\mu$ have been limited to
small $\mu$ region due to the complex fermion determinant,
known as the ``sign problem'' \cite{muroya03}. One of several methods circumventing this
problem is to use an imaginary chemical potential $\mu=i\mu_I$.
Indeed, this method has provided transition lines in the $T-\mu$ plane via
an analytic continuation from those obtained at imaginary $\mu$
\cite{forcrand02:_qcd,d'elia-lombardo,d'elia07}. Moreover, it has been known that there is a
phase transition specific to the imaginary chemical potential
characterizing the deconfinement phase at high temperature
\cite{roberge86:_gauge_qcd}. Rich phase structures later found in the
lattice simulations provide a testing ground for understanding the
nature of phase transitions in QCD
\cite{forcrand:_const_qcd,bonati:_rober_weiss_endpoin_in_n_f_qcd,Chen, Nagata}.
Those properties give constraints on model studies which can be extended
to real $\mu$. In this work, we study the phase structure of the
Polyakov-loop extended Nambu-Jona-Lasinio (PNJL) model
\cite{fukushima04:_chiral_polyak,ratti06:_phases_qcd} which satisfies
fundamental symmetries of QCD relevant for phase transitions at
imaginary chemical potential. Focusing on the deconfinement transition,
we show that the ``statistical confinement'' feature of the model
naturally leads to characteristic behaviors of the order parameters while details
depend on the choice of the Polyakov loop potential. We discuss dual
parameters to characterize the phase transitions. Finally, we point out the
existence of the critical endpoint (CEP) associated with the
deconfinement transition at imaginary chemical potential and clarify the
relation between its location and the chiral phase transition.
In the next section, we will give a brief introduction of the model. We
will discuss the characteristic behavior of the order parameters as well
as the dual parameters in Sec.~\ref{sec:orderparameter}. The critical
endpoint of the deconfinement transition will be discussed in
Sec.~\ref{sec:cep} and Section \ref{sec:summary} is devoted to the
summary. More details can be found in Ref.~\cite{morita11:_probin_decon_in_chiral_effec}.
\section{PNJL model at imaginary chemical potential}
The Lagrangian of the two-flavor PNJL model is given by
\begin{equation}
\mathcal{L} = \bar{q}(i \gamma_\mu D^\mu -m_0 )q +
G_s[(\bar{q}q)^2+(\bar{q}i\gamma_5 \vec{\tau}q)^2]-\mathcal{U}(\Phi[A],\Phi^*[A];T).
\end{equation}
The model is an extension of the NJL model, which is an
effective model of chiral properties of QCD
\cite{nambu61:_NJLI,hatsuda94:_qcd_lagran}, such that
quarks couple with background gluonic fields described by a $Z(3)$
symmetric effective potential $\mathcal{U}$ which takes care of
confinement.
In the covariant derivative
$D^\mu = \partial^\mu - i A^\mu$, only the temporal components of
$A_0= gA_0^a \lambda^a/2$ is included. The effective potential
$\mathcal{U}$ is expressed in terms of the traced Polyakov loop and its conjugate,
$\Phi = \langle {\rm Tr}_c L \rangle/3$ and
$\Phi^* = \langle {\rm Tr}_c L^\dagger \rangle/3$, respectively.
This coupling between quarks and gluons leads to an almost simultaneous
crossover of the chiral and deconfinement transitions at finite
temperature, of which order parameters are chiral condensate
$\sigma\equiv\langle\bar{q}q\rangle$ and the Polyakov loop $\Phi$
\cite{fukushima04:_chiral_polyak}, provided the Polyakov loop potential
$\mathcal{U}$ yields a first order transition at $T_0=270$ MeV in
accordance with pure $SU(3)$ lattice calculations.
Two functional forms of $\mathcal{U}$, which reproduce the thermodynamic quantities
obtained in pure $SU(3)$ lattice gauge theory \cite{Boyd}, have been used.
One has a polynomial form
\begin{equation}
\frac{\mathcal{U}_{\rm poly}}{T^4} = -\frac{b_2(T)}{2}\Phi^* \Phi
-\frac{b_3}{6}[\Phi^3+(\Phi^*)^3]+\frac{b_4}{4}(\Phi^* \Phi)^4\label{eq:upol}
\end{equation}
with a set of parameters given in \cite{ratti06:_phases_qcd}.
The other is a logarithmic one \cite{roessner07:_polyak}
\begin{equation}
\frac{\mathcal{U}_{\rm log}}{T^4} = -\frac{a(T)}{2}\Phi^* \Phi +
b(T)\log\{1-6\Phi^* \Phi+4[\Phi^3 + (\Phi^*)^3]-3(\Phi^* \Phi)^2\}.\label{eq:ulog}
\end{equation}
The logarithm restricts possible values of $\Phi$ and $\Phi^*$ to the
so-called target space, since the argument of the logarithm must be
positive.
At imaginary $\mu$, the two Polyakov loop variables $\Phi$
and $\Phi^*$ are complex conjugate
\cite{sakai08:_polyak_nambu_jona_lasin}.
Moreover, the partition function of the PNJL model at imaginary chemical
potential has been shown \cite{sakai08:_polyak_nambu_jona_lasin} to
have the same periodicity in $\theta=\mu_I/T$ as that of QCD, $Z(\theta+2\pi/3)=Z(\theta)$,
which was pointed out by Roberge and Weiss \cite{roberge86:_gauge_qcd} as a
remnant of $Z(3)$ symmetry.
Therefore we may express them by using a modulus and a phase
$\Phi = |\Phi|e^{i\phi}$ and $\Phi^* = |\Phi|e^{-i\phi}$.
The thermodynamic potential in the mean field approximation reads
\begin{eqnarray}
\Omega(T,V,\theta) &=& (G_s \sigma^2 + \mathcal{U})V -4V \int\frac{d^3 p}{(2\pi)^3}
\left[ 3(E_p-E_p^0) \right.\nonumber \\
+T\ln[\!\!\!&1&\!\!\!+3|\Phi|e^{i(\theta+\phi)-\beta E_p} +3 |\Phi|
e^{i(2\theta-\phi)-2\beta E_p}+e^{3i\theta-3\beta E_p}] \nonumber \\
+T\ln[\!\!\!&1&\!\!\!+3|\Phi|e^{-i(\theta+\phi)-\beta E_p} +3 |\Phi|
e^{i(\phi-2\theta)-2\beta E_p}+e^{-3i\theta-3\beta E_p}]]\label{eq:omega}
\end{eqnarray}
where $E_p=\sqrt{p^2+M^2}$, $E_p^0=\sqrt{p^2+m_0^2}$, and
$M=m_0-2G_s\sigma$. The first term in the momentum integral is a
divergent vacuum term, which is regularized by a three-momentum cutoff
$\Lambda$. The cufoff and coupling are fixed to $G_s=5.498$ GeV$^{-2}$
and $\Lambda=0.6315$ GeV so as to reproduce the vacuum pion mass and
pion decay constant with $m_0=5.5$ MeV. In the following we mainly focus
on the result in the chiral limit $m_0=0$ to preserve the chiral symmetry in
the Lagrangian.
The chiral condensate $\sigma$ serves as an order
parameter for the chiral phase transition. The order
parameters are determined by the minimum of the potential which is
obtained by solving the gap equation $\partial \Omega/\partial X_i=0$ with
$X_i = M, |\Phi|, \phi$.
\section{Behavior of the order parameters}
\label{sec:orderparameter}
\subsection{Order parameters at imaginary chemical potential}
\begin{figure}[!t]
\epsfig{file=PNJL_Sigma-theta_1x1_m0mono.eps,width=0.45\textwidth}
\epsfig{file=PNJL_Phi-theta_m0mono.eps,width=0.45\textwidth}
\caption{Left : Chiral condensate for various temperatures as functions
of $\theta$. Right : Phase of the Polyakov loop $\phi$. Both results
are in the chiral limit $m_0=0$ and for the logarithmic potential.}
\label{fig:sigma}
\epsfig{file=PNJL_Poly-theta_m0mono.eps,width=0.45\textwidth}
\epsfig{file=logpotential_target.eps,width=0.4\textwidth}
\caption{Left : Modulus of the Polyakov loop $|\Phi|$. Right : target
space of the Polyakov loop on complex $\Phi$ plane. A region inside the
solid lines denote the target space of the logarithmic potential in
which the argument of the logarithm is positive.}
\label{fig:moduls_target}
\end{figure}
First we consider two extreme limits in order to see characteristic
$\theta$ dependences of $\sigma$ which has the same periodicity $2\pi/3$
as $\Omega$.
Expanding Eq.~(\ref{eq:omega}) for small $e^{-\beta E_p}$, we have a gap
equation at small $|\Phi|$ limit in which only a term proportional to
$\cos3\theta$ remains with a small magnitude $\sim e^{-3 \beta E_p}$
indicating the statistical confinement.
This dependence naturally leads to the periodicity
$2\pi/3$ when $|\Phi|$ is negligible and chiral symmetry is broken.
On the other hand, when $|\Phi|\simeq 1$, the model reduces to the NJL
model except for the coupling of $\phi$ with $\theta$, as seen in
Eq.~(\ref{eq:omega}). In this case, the apparent $\theta$ dependence is governed by
$\cos\theta$ as a consequence of deconfinement. Although this factor
does not match with the required periodicity $2\pi/3$, it is preserved
by a change of $\phi$, namely, the Roberge-Weiss transition.
We show the chiral condensate in the left panel of
Fig.~\ref{fig:sigma} obtained by numerically solving the gap equations.
One sees that $\sigma$ at low temperature ($T=220$ MeV) exhibits small and smooth
variation as a function of $\theta$, as discussed above. On the other
hand, one sees a cusp at $\theta=\pi/3$ and $T=280$ MeV.
This is a consequence of a Roberge-Weiss transition depicted in
the right of Fig.~\ref{fig:sigma}, in which the phase $\phi$ changes from 0 to
$-2\pi/3$, smoothly at low $T$ but discontinuously at high $T$.
As a result, the required periodicity of $\sigma$ is preserved.
One also sees a second order chiral phase transition for $T=280$ MeV
in which the chiral symmetry is broken around $\theta=\pi/3$. This implies the
chiral critical temperature at imaginary $\mu$ is higher
than that of zero and real $\mu$. This can be also
understood from the gap equation for $\sigma$, since $\cos n\theta$ is
replaced by $\cosh n \beta \mu$ for real $\mu$.
While the above properties are independnent of the choice of
$\mathcal{U}$, there are some potential dependent features as follows.
In Fig.~\ref{fig:sigma} and the left of Fig.~\ref{fig:moduls_target},
one sees a discontinuity in the order parameters at the same $\theta$.
This shows a first order deconfinement transition which exists only in
the case of the logarithmic potential (\ref{eq:ulog}).
The polynomial potential (\ref{eq:upol}) exhibits smoother change
near phase transition.
In the right of Fig.~\ref{fig:moduls_target}, the target space of the
Polyakov loop is displayed. Owing to the $Z(3)$ symmetry,
$\mathcal{U}$ has three degenerate minima at $T > T_0$. Putting quarks
into the system makes one of those minima favored. While ${\rm Im}\Phi=0$
is always chosen at $\theta=0$, ${\rm Im}\Phi \neq 0$ is favored at
imaginary chemical potential due to the coupling of $\theta$ and $\phi$
seen in Eq.~(\ref{eq:omega}). At low temperature where minimum of
$\mathcal{U}$ is close to the origin, the minimum of the effective potential
smoothly moves from $\phi=0$ to $\phi=-2\pi/3$ across $\theta=\pi/3$.
At high temperature, however,
there is a potential wall which makes the transition from point A to
point B discontinuous. Since the polynomial potential does not have any
restriction of the target space, the minimum passes outside (C) the target space
near the RW transition.
\begin{figure}[t]
\epsfig{file=pd-th_log.eps,width=0.45\textwidth}
\epsfig{file=pd-th_pol.eps,width=0.45\textwidth}
\caption{Phase diagram on $T-\theta$ plane. Solid lines, dotted lines,
and dashed lines stand for first, second, and crossover transitions,
respectively.}
\label{fig:pd}
\end{figure}
The phase diagrams shown in Fig.~\ref{fig:pd} summarize the behavior of the
order parameters. One sees a first order deconfinement transition
and an associated critical endpoint (CEP) only for the logarithmic
potential. This also implies that the RW endpoint, where the first order RW
transition terminates, is a triple point. On the other hand, one sees a
second order RW endpoint for
the polynomial potential. The properties of the RW endpoint in QCD might reflect
the nature of the QCD phase transition at real $\mu$
We refer to Refs.~\cite{bonati:_rober_weiss_endpoin_in_n_f_qcd}
and \cite{forcrand:_const_qcd} for recent calculations of $N_f=2$ and $N_f=3$ latttice QCD,
respectively. Especially it should be noted that the order of the RW
endpoint has a non-trivial bare quark mass dependence which cannot be
reproduced by chiral effective models. (See Sec.~\ref{sec:cep})
An improved model was proposed in Ref.~\cite{sakai:_10063408} to
reproduce this property.
\subsection{Dual parameters for deconfinement}
\begin{figure}[t]
\epsfig{file=PNJL_dual-Poly_2x2mono.eps,width=0.9\textwidth}
\caption{Dual parameters compared with Polyakov loop. Top panels show
$\Phi$, $\Sigma^{(1)}$ and $\Sigma_\theta^{(1)}$ for
$\mathcal{U}_{\rm log}$ (left) and $\mathcal{U}_{\rm poly}$ (right)
while bottom ones displays their derivatives with respect to temperature.}
\label{fig:dual}
\end{figure}
It has been shown that information on the deconfinement is
encoded in $\theta$ dependence of $\sigma$. We can consider dual
parameters which characterize the deconfinement transition.
A dual parameter was introduced in \cite{bilgici08:_dual_polyak}.
By considering a twisted boundary condition for quarks
$q({\mathbf{x}},\beta)=e^{i\varphi}q(\mathbf{x},0)$, one may define the
corresponding chiral condensate $\sigma(\varphi)$. Then the dual chiral
condensate $\Sigma^{(n)}$ reads
\begin{equation}
\Sigma^{(n)}(T) =
-\int_{0}^{2\pi}\frac{d\varphi}{2\pi}e^{-in\varphi}\left[
-\frac{1}{V}\left\langle{\rm
Tr}[(m_0+D_\varphi)^{-1}]\right\rangle
\right]
\end{equation}
While the twisted boundary condition is similar to introducing imaginary
chemical potential \cite{weiss87:_how}, it does not apply to the
background gauge field. Therefore, $\sigma(\varphi)$ has a periodicity
$2\pi$ and was calculated in a PNJL model by fixing the Polyakov loop at
its $\theta=0$ value \cite{kashiwa09:_dual_polyak_nambu_jona_lasin}.
Particularly $\Sigma^{(1)}$ is called dressed Polyakov loop, since it
has the same transformation properties under $Z(3)$ and thus is expected
to serve as an order parameter of the deconfinement transition.
Analogously, we consider a modified dual parameter which utilizes the characteristic
property of $\sigma(\theta)$,
\begin{equation}
\Sigma_\theta^{(n)}(T) = \frac{3}{2\pi}\int_{-\pi/3}^{\pi/3}d\theta e^{-in\theta}\sigma(T,\theta).
\end{equation}
where we take the integration range $[-\pi/3,\pi/3]$, owing to the
periodicity of $\sigma(\theta)$.
We compare those dual parameters for $n=1$ with the Polyakov loop in
Fig.~\ref{fig:dual}, as well as their derivatives with respect to
temperature, of which peaks can be regarded as (pseudo)critical
temperatures. One sees that while dual parameters show a rapid increase
as seen in the Polyakov loop (top),\footnote{Dual parameters are normalized to
0 as $T\rightarrow 0$ and 1 as $T\rightarrow\infty$
\cite{morita11:_probin_decon_in_chiral_effec}.} their derivatives
exhibit different peak structures. The derivatives of the dual parameters have a peak at
the chiral transition temperature, independent of $\mathcal{U}$.
As for the deconfinement, however, existence of the peak depends on
$\mathcal{U}$.
The dressed Polyakov loop exhibits a peak for the
$\mathcal{U}_{\rm log}$ for which $|\Phi|$ shows stronger crossover than
$\mathcal{U}_{\rm poly}$. Moreover, the modified dual parameter exhibits
only a shoulder even for $\mathcal{U}_{\rm log}$. This result indicates
different sensitivity of the dual parameters to the chiral and
deconfinement transition.
\section{Critical endpoint of deconfinement}
\label{sec:cep}
\begin{figure}[!t]
\epsfig{file=gs-mu2mono.eps,width=0.45\textwidth}
\epsfig{file=pd-gs6500mono.eps,width=0.45\textwidth}
\caption{Left : location of the deconfinement CEP
$(T_{\rm CEP}, \mu^2_{\rm CEP})$ as a function of $G_s$. Right : Phase
diagram for $G_s=6.5$ GeV$^{-2}$ which gives $\mu_{\rm CEP}^2 > 0$.}
\label{fig:deccep}
\end{figure}
Now let us turn to the deconfinement CEP found in the case of
$\mathcal{U}_{\rm log}$. Here we vary the four-fermion coupling constant
$G_s$ to preserve the chiral symmetry in the Lagrangian. Locations of the CEP are shown in
the left of Fig.~\ref{fig:deccep} for various values of $G_s$. One sees the squared critical chemical
potential $\mu_{\rm CEP}^2$ increases with $G_s$ to reach
$\mu_{\rm CEP}^2=0$ around $G_s\simeq 6.3$ GeV$^{-2}$.
In the right of Fig.~\ref{fig:deccep}, we also depict a phase diagram
for $G_s=6.5$ GeV$^{-2}$ in which the CEP exists at real chemical
potential. One sees that the first order deconfinement transition starting
from the RW endpoint (see Fig.~\ref{fig:pd}) is prolonged, while the
chiral critical line moves upward. The relation between these two
changes can be understood as follows. Since the Polyakov loop potential
$\mathcal{U}_{\rm log}$ has a first order phase transition at $T=T_0$, the model
results in the same transition when the effects of quarks are negligible
in thermodynamics. The contribution of quarks to thermodynamic potential
is essentially determined by the dynamical quark mass $M=m_0-2G_s\sigma$, not by the
current quark mass $m_0$, as seen in Eq.~(\ref{eq:omega}). When
dynamical quark mass becomes lighter around $T=T_0$, the deconfinement transition is
modified to a crossover one.
As $G_s$ increases, the stronger coupling leads to a larger condensate
$|\sigma(T=0)|$ thus the dynamical quark mass becomes heavier. This appears as the modified chiral critical line in
the phase diagram at $G_s=6.5$ GeV$^{-2}$ and the resultant dynamical
quark mass is heavy enough to recover the first order the deconfinement
transition. At the reference value of $G_s$, the imaginary chemical
potential
weakens the thermal terms by $\cos n\theta$ in the thermodynamic potential thus resembling a
heavier quark mass which yields the CEP and a first order transition.
While the above consideration is completely independent of the form of
$\mathcal{U}$, quantitative features such as the value of dynamical quark
mass which makes the transition first order depend on the choice of $\mathcal{U}$.
\begin{figure}[!t]
\epsfig{file=PNJL_Poly-T-Gs_Logmono.eps,width=0.45\textwidth}
\epsfig{file=PNJL_Poly-T-largeGs_polmono.eps,width=0.45\textwidth}
\caption{Behavior of $|\Phi|$ for large $G_s$. Left :
$\mathcal{U}_{\rm log}$. Right : $\mathcal{U}_{\rm poly}$}
\label{fig:polgs}
\end{figure}
Figure \ref{fig:polgs} shows the behavior of $|\Phi|$ for various $G_s$ at vanishing
chemical potential. One sees that $|\Phi|$ becomes steeper for larger
$G_s$ in both of $\mathcal{U}$.
The case of $\mathcal{U}_{\rm log}$ has a discontinuity already at
$G_s=6.5$ GeV$^{-2}$ as mentioned above. $\mathcal{U}_{\rm pol}$, which has a smoother variation of $|\Phi|$
against $T$, eventually approaches the pure gauge case for much
larger $G_s$. At $G_s=25$ GeV$^{-2}$,
where the dynamical quark mass at $T=0$ is around 2.5 GeV, the first
order deconfinement transition is recovered. The origin of this difference
is the much weaker first order transition in $\mathcal{U}_{\rm poly}$,
which easily turns into crossover when quarks heavier than 2.5 GeV are
put into the system. If one characterizes a strength of the
deconfinement transition by a gap of the Polyakov loop $\Delta \Phi$ at $T=T_0$,
one finds $\Delta \Phi=0.47$ for $\mathcal{U}_{\rm log}$ and 0.072 for
$\mathcal{U}_{\rm poly}$. Since $G_s$ determines the scale of the
dynamical chiral symmetry breaking, $\sigma(T=0)$, this result indicates
an interplay of the two transitions which have a unique scale,
$\Lambda_{\rm QCD}$, in the case of QCD.
It has been shown that a first order deconfinement phase transition also emerges in the large
$N_c$ limit of the PNJL model
\cite{mclerran09:_quark_matter_and_chiral_symmet_break}. This result
has a common origin with the present study in a sense that taking large
$N_c$ limit makes the system gluon dominated due to $1/N_c$ suppression of the quark
contribution, while large $G_s$ thermally suppresses quarks in
the chirally broken phase. In our case, however, chiral transition
temperature moves upward thus there is a discrepancy between the
deconfinement and chiral transition temperatures, in contrast to the
large $N_c$ limit with a fixed $G_s N_c$ in \cite{mclerran09:_quark_matter_and_chiral_symmet_break}.
\section{Summary}
\label{sec:summary}
We have explored the deconfinement transition in the PNJL model at imaginary
chemical potential. We point out that the chiral condensate at imaginary
chemical potential, $\sigma(\theta)$, has a characteristic $\theta$
dependence due to the deconfinement property which naturally arises from the
statistical confinement feature of the model. While the confined
phase is characterized by a smooth $\cos3\theta$ dependence, the
deconfined phase exhibits $\cos\theta$ dependence together with cusps at
$\theta=\pi/3$ (mod $2\pi/3$) induced by the abrupt change of the phase of
the Polyakov loop (Roberge-Weiss transition).
We introduce a new dual parameter utilizing this $\theta$ dependence and
compare it with the Polyakov loop and the dressed Polyakov loop.
Different sensitivities of these parameters to chiral and
deconfinement transitions are found.
Changing the four fermion coupling constant, we found that an
interplay between the thermal quark contribution through the dynamical
chiral symmetry breaking and the Polyakov loop potential
determines the location of the deconfinement CEP at imaginary chemical
potential. In particular, we found that the deconfinement CEP can be located
in the real chemical potential regime for a Polyakov loop potential with a strong
first order transition and large dynamical chiral symmetry breaking.
We expect that these results are useful for understanding of the QCD
phase transition.
This work is supported by the Yukawa International Program for
Quark-Hadron Sciences at Kyoto University. K.M. and V.S. acknowledges
FIAS for support. B.F. and K.R. acknowledges partial support by
EMMI. K.R. acknowledges partial support by the Polish Ministry of
Science (MEN).V.S. was supported by the U.S. Department of Energy under
Contract No. DE-AC02-98CH10886.
|
\section{Introduction}
Nonlinear Schr\"{o}dinger equation (NLSE) describes a broad range of physical phenomena, e.g. nonlinear modulation of collisionless plasma waves \cite{Ablowitz}, self trapping of a light beam in a color dispersive system \cite{Agarwal}, helical motion in a very thin vortex filament \cite{Hasegawa}, propagation of heat pulses in an-harmonic crystals \cite{Drazin}, modulation instability in water waves \cite{Hasegawa}, etc. In optical fibers, the soliton solutions of the NLSE provide a secure means to carry bits of information over many thousands of miles \cite{Agarwal}. Termed as the Gross-Pitaveskii equation, the NLSE with an appropriate potential can be utilized to describe the dynamics of the Bose-Einstein condensate, both with the attractive and repulsive nonlinearities \cite{Hasegawa, Malomed}. It is our objective in this paper to solve the Gross-Pitaveskii equation equipped with a point-like potential to find the critical values of the soliton velocities when the amplitude of the point-like potential is either very small $(\sim 10^{-1} - 10^{-2})$ or large $(\sim 2.5 - 4.5)$ compared to the soliton amplitude which is the unity in our paper.
We know the soliton solution of the homogeneous NLSE
\begin{eqnarray}
i\partial_{t}u + \frac{1}{2}\partial^{2}_{x}u + u|u|^{2} = 0, \quad -\infty < x< \infty, \; t > 0 ,
\label{NLSE}
\end{eqnarray}
with initial condition $u_0$ given by
\begin{equation}
u_{0}(x) = A \mathrm{sech}\left(A\left(x\right)\right)e^{\left(i\phi +iVx\right)}
\label{initial0}
\end{equation}
is given by
\begin{eqnarray}
u(x,t) = A \mathrm{sech} \left(A\left(x-Vt\right)\right)\exp\left(i\phi + iVx + \frac{i}{2} \left(A^{2}-V^{2}\right)t\right), \; A > 0, \; V \in \mathbb{R},
\label{NLSsol}
\end{eqnarray}
where $A$ is the soliton amplitude, $V$ the soliton velocity, and $\phi$ the phase lag.
Consider a perturbed NLSE, that is, the Gross-Pitaveskii equation by adding an external potential, $-\epsilon \delta(x)u$,
\begin{eqnarray}
\left\{ \begin{array}{cc} i\partial_{t}u + \frac{1}{2}\partial^{2}_{x}u + u|u|^{2} = -\epsilon \delta(x)u, \\
u(x,0) = u_{0}(x), \end{array} \right.
\label{GP1}
\end{eqnarray}
where $\delta(x)$ is the Dirac delta function with a constant $\epsilon \in \mathbb{R}$. Such an external potential represents the impurity or defect in the optical fiber. The well-posedness of the equation
\begin{eqnarray}
\left\{ \begin{array}{cc} i\partial_{t}u + \frac{1}{2}\partial^{2}_{x}u + u|u|^{p-1} = -\epsilon \delta(x)u, \\
u(x,0) = u_{0}(x), \end{array} \right.
\label{GP2}
\end{eqnarray}
with $p\geq 1$ and initial data $u_0$ in $H^1({\ensuremath{\mathbb{R}}})$, has been extensively studied and is based on the knowledge of the self-adjoint (in $L^2$) operator $-\partial_{xx}+\epsilon\delta$. Using \cite{Cazenave}, Le Coz {\it et al.} proved the existence of a time $T>0$ and of a unique solution to Eq. \ref{GP1} (where $\epsilon \in {\ensuremath{\mathbb{R}}}$) in $C\big([0,T),H^1({\ensuremath{\mathbb{R}}})\big)\cap C^1\big([0,T),H^{-1}({\ensuremath{\mathbb{R}}})\big)$ satisfying $\lim_{t \rightarrow T} \|\partial_xu\|_2=\infty$. Moreover the energy is conserved in time. This result was extended to $p\geq 1$ by Fukuizumi {\it et al.} in \cite{Fuku2}. For $p=3$ (more generally $p\in (1,5)$), global existence in $H^1$ also holds by Gagliardo-Nirenberg's inequality and energy conservation. Global existence in $H^1$, is also discussed by Goodman {\it et al.} \cite{GHW} using a fixed-point argument and time-invariance of the $L^2$-norm and of the Hamiltonian derived from the NLSE. Notice that the study of stability of nonlinear bound states which are solutions of the form $\exp(-{\tt i}\omega t)\phi_{\omega}(x)$ with $\omega>0$, and for which:
\begin{eqnarray}
\label{NLSB}
-\frac{1}{2}\partial_{xx}\phi_{\omega} -\epsilon \phi_{\omega} - |\phi_{\omega}|^2 \phi_{\omega}=\omega \phi_{\omega}
\end{eqnarray}
plays an important role in the theory of NLSE with defect and could possibly be useful numerically. Explicit formulas and stability analysis for $\phi_{\omega}$ can be found in \cite{Fuku2,LeCoz}.
If we now take a soliton approaching the impurity from the left as an initial condition $u_{0}$:
\begin{equation}
u_{0}(x) = A \mathrm{sech}\left(A\left(x-x_{0}\right)\right)e^{\left(i\phi +iVx\right)},\; x_{0} \ll 0,
\label{initial1}
\end{equation}
then until the time $t_{0} = \frac{x_{0}}{V}$, the solution will still be given by
Eq. \ref{NLSsol}. In this paper we consider $A = 1$ and $\phi = 0$. Thus the soliton velocity $V$ and the strength of the impurity $\epsilon$ are the only parameters of the problem.
For $t_{0} > \frac{x_{0}}{V}$, the effects of the potential are highly visible and a lot of research has been done on the transmission and reflection coefficients of the $\delta$-potential by the standard scattering theory \cite{Holmer}.
Malomed and his co-workers \cite{Cao_Malomed, Malomed} showed mainly numerically, that for any given velocity $V \left( > 0 \right)$, there exists a threshold value $\epsilon_{thr} \left( > 0 \right)$ of $\epsilon$, for which the soliton can marginally pass through the defect. So for the given velocity $V$, if $\epsilon < \epsilon_{thr}$, the soliton can pass through the defect and the soliton gets trapped otherwise. They considered the soliton-soliton collisions within the coupled NLSE. In the limiting condition one soliton has very large amplitude and is very narrow accordingly, while the soliton governed by the other equation has finite amplitude and width. In this limiting condition the two coupled NLSE are reduced to a single equation, in which the narrow soliton will be represented by the $\delta$-function,
\begin{equation}
i\partial_{t}u + \frac{1}{2}\partial^{2}_{x}u + u|u|^{2} = -\epsilon \delta(x)u. \nonumber
\label{NLSE2}
\end{equation}
Holmer and his co-workers studied the NLSE with $V \gg 1$ \cite{Holmer1} and $V \ll 1,\; \epsilon \ll 1$ \cite{Holmer2}. They showed for high $V$, there exits the bound state which is given by
$$
u(x,t) = e^{i\lambda^{2}\frac{t}{2}}\lambda \mathrm{sech}\left(\lambda |x| + \tanh^{-1}\left(\epsilon/\lambda\right)\right), \quad 0<\lambda <\epsilon,
$$
and this bound state is ``left behind" after the interaction (see bottom right figure of Figure \ref{fig:soliton3}). Also they proved in \cite{Holmer2} that for $V \ll 1$ and $ \epsilon \ll 1$, the solution can be approximated by the soliton solution of the homogeneous NLSE $\left(\epsilon = 0\right)$.
To solve Eq. \ref{NLSE2} for any given $\epsilon \left( > 0 \right)$ and $V \left( > 0 \right)$, we consider three cases: (a) small value of $\epsilon, \; \mathrm{where} \; \epsilon \le 0.3$ (b) moderate value of $\epsilon, \; \mathrm{where} \; 0.3 < \epsilon \le 3.5$ \cite{Cao_Malomed} and (c) large value of $\epsilon, \; \mathrm{where} \; \epsilon > 3.5$. For solving Eq. \ref{NLSE2}, one can use the Split Step Fourier Method (SSFM) to reduce the computational time. To get $\epsilon_{thr}$ for any given $V$ with certain accuracy one must conduct a series of simulations. The number of simulations increase with the increase of the level of accuracy. In addition to conduct a series of simulations with small time steps, one needs a large amount of the computational time. This is our main motivation to propose a suitable method to overcome such a high computational complexity by using the generalized polynomial chaos (gPC) methods \cite{XiuBook}.
The gPC method belongs to the class of non-sampling methods \cite{Xiu2, Xiu3}. In this method the stochastic quantities are expanded by orthogonal polynomials. Different types of orthogonal polynomials can be chosen for better convergence. The gPC expansion is a spectral representation in random space and exhibits fast convergence when the expanded function depends smoothly on the random parameters \cite{Gottlieb}. When the gPC method is applied to solve any differential equation, the main computational work is needed to solve the expansion coefficients of the gPC expansion. A common approach is the Galerkin method that minimizes the residue in the polynomial space. The stochastic Galerkin (SG) approach, however, would be extremely difficult to use when the governing stochastic equations take complicated forms. In our case, the NLSE contains the nonlinear term $|u|^{2}u$. For the SG method, it is very hard to get the corresponding explicit deterministic equations after expanding the nonlinear terms. So that, in this work we use the high-order stochastic collocation (SC) approach \cite{Xiu2} that combines the advantages of both the Monte Carlo sampling and the gPC-Galerkin methods. The gPC method reduces the number of simulations for finding the critical velocity, $V_{c}$, for any given value of $\epsilon$ thanks to the high-order convergence of the method. Since the equation has only two parameters, i.e. $\epsilon$ and $V$, we treat at least one of them as a stochastic variable in the gPC framework. In the present work we consider $V$ as the stochastic variable and let $\epsilon$ be fixed. So for any given $\epsilon$, we find $V_{c}$, the critical value of $V$ around which the soliton is either transmitted or trapped. Thus it is obvious that for $V > V_{c}$, the soliton passes through the defect. By adopting this idea we develop a step-by-step gPC collocation method to find the critical velocity of the soliton.
In \cite{Cao_Malomed} the relation between $\epsilon_{thr}$ and $V$ was obtained only for the moderate values of $\epsilon$, i.e. for those comparable to the soliton amplitude $A = 1$. But the results of the numerical simulations for very small or large values of $V$ were not obtained, perhaps due to the huge computational burden. By the gPC method, we were able to reduce the overhead computational time, for having detailed simulations performed for large and small values of $\epsilon$ to find the corresponding critical velocity $V_{c}$.
Since the analysis for the moderate values of $\epsilon$ are already done \cite{Cao_Malomed}, we do not intended to repeat the analysis for those values of $\epsilon$ in this paper. Here we mainly focus on the small and high values of $\epsilon$. For the small values of $\epsilon$, the gPC takes much longer time than the gPC method for the large values of $\epsilon$ due to the extremely small critical velocities.
This paper is organized as follows. In Section $2$, we discuss the SSFM. Section $3$ describes the gPC collocation method. Section $4$ contains the gPC collocation algorithm for the NLSE with the singular potential term to detect the critical velocity for the given value of $\epsilon$. Section $5$ presents the numerical results. Concluding remarks and future works are presented in Section $6$.
\section{Split Step Fourier Method}
The SSFM is a pseudo-spectral numerical method used to solve nonlinear PDEs like the NLSE. Eq. \ref{NLSE} can be rewritten as
\begin{eqnarray}
\frac{\partial u}{\partial t} = i\left[N + D\right]u,
\label{split1}
\end{eqnarray}
where $D = \frac{1}{2}\frac{\partial^{2}}{\partial x^{2}}$ and $N = |u|^{2}$.
The solution of Eq. \ref{split1} can be written as
$$
u(x,t) = e^{it\left(D+N\right)}u(x,0),
$$
where $u(x,0)$ is the initial condition. Since $D$ and $N$ are the operators, they do not necessarily commute. However the Baker-Hausdorff formula can be applied to show that the error will be of order $dt^{2}$ if we are taking a small but finite time step $dt$ \cite{sinkin}. We therefore can write
\begin{equation}
u\left(x,t+dt\right) \approx e^{idtN}e^{idtD}u(x,t).
\label{split_order1}
\end{equation}
The part of this equation involving $N$ can be computed directly using the wave function $u(x,t)$ at time $t$. To compute the exponential involving $D$ we use the fact that in the frequency domain, the partial derivative operator $\frac{\partial}{\partial x}$ is converted into $ik$, where $k$ is the frequency associated with the Fourier transform. Then we take the Fourier transform of $u(x,t)$ recover the associate wave number, and compute
$$
e^{-\frac{1}{2}idtk^{2}}\mathbb{F}\left[u(x,t)\right],
$$
where $\mathbb{F}$ denotes the Fourier transform. Then we take the inverse Fourier transform of the expression to find the solution in the physical space, yielding the final expression
$$
u(x, t+dt) = e^{idtN}\mathbb{F}^{-1}\left(e^{-\frac{1}{2} idtk^{2}} \mathbb{F} \left[u(x,t)\right]\right).
$$
We apply SSFM to Eq. \ref{NLSE2} where the nonlinear operator $N = |u|^{2}$ and the linear operator $L = \frac{1}{2}\frac{\partial^{2} }{\partial x^{2}} + \epsilon \delta(x)$.
In our numerical simulations we use the high-order SSFM, such as the Strang splitting based on:
\begin{eqnarray*}
e^{i\left(L+N\right)\Delta t} & = & e^{i L\frac{\Delta t}{2}}e^{i N \Delta t}e^{i L\frac{\Delta t}{2}}+\mathcal{O}\big(\Delta t^3([L,[L,N]]+[N,[N,L]])\big).
\end{eqnarray*}
where $[L,N]=LN-NL$ denotes the commutator between $L$ and $N$. Thus, from $t$ to $t+\Delta t$
\begin{eqnarray}
u(x,t+\Delta t) & = & e^{i\left(L+N\right)\Delta t}u(x,t), \nonumber \\
&\approx & e^{i L\frac{\Delta t}{2}}e^{i N \Delta t}e^{i L\frac{\Delta t}{2}}u(x,t).
\label{split_order2}
\end{eqnarray}
\section{gPC collocation method}
We solve Eq. \ref{GP1} with the initial condition given by Eq. \ref{initial1} for both small and large values of $\epsilon$ by the gPC collocation method.
We use the gPC method for the solution of the NLSE using the Wiener-Askey scheme \cite{XiuBook, Xiu3}, in which Hermite, Legendre, Laguerre, Jacobi and generalized Laguerre orthogonal polynomials are used for modeling the effect of continuous random variables described by the normal, uniform, exponential, beta and gamma probability distribution functions (PDFs), respectively \cite{Chakraborty_Jung, Xiu2}. These orthogonal polynomials are optimal for those PDFs since the weight function in the inner product and its support range correspond to the PDFs for those continuous distributions.
Following the standard gPC expansion, we assume that $u(x,t, \xi)$ is sufficiently smooth in $\xi$ and has a converging expansion of the form
$$u(x,t,\xi) = \sum_{k=0}^{\infty}\hat{u}_{k}(x,t)P_{k}(\xi),$$
where the orthonormal polynomials $P_{k}(\xi)$ correspond to the PDF of the random variable $\xi$ and satisfy the following orthogonality relation:
$$
\mathbf{E}[P_{k}P_{l}] := \int P_{k}(\xi)P_{l}(\xi)\rho(\xi)d\xi = \delta_{kl}.
$$
Here $\delta_{kl}$ is the Kronecker delta and $\rho(\xi)$ is the weight function. Note that the polynomials are normalized.
For the stochastic collocational approach we approximate $\hat{u}_{k}(x,t)$ as,
\begin{equation}
\hat{u}_{k}(x,t) = \sum_{j=0}^{Q}u\left(x,t,p^{j}\right)P_{k}\left(p^{j}\right)\alpha_{j}, \; k = 0, \cdots, Q,
\end{equation}
where $Q+1$ is the total number of the collocation nodes.
Here $\left\lbrace p^{j},\alpha^{j} \right\rbrace$ is a set of nodes and weights, where $p^{j}$ and $\alpha^{j}$ denote the $j$-th node and its associated weights, respectively, in the random space $\Gamma$ such that
\begin{equation}
\mathbb{W}^{Q}\left[f\right]\equiv \sum_{j=0}^{Q}f\left(p^{j}\right)\alpha^{j},
\end{equation}
is an approximation of the integral
\begin{equation}
\mathit{I}\left[f\right]\equiv \int_{\Gamma}f(p)\rho(p)dp = \mathbb{E}\left[f(p)\right],
\end{equation}
for sufficiently smooth functions $f(p)$, i.e,
$$
\mathbb{W}^{Q}\left[f\right] \rightarrow \mathit{I}\left[f\right], \; Q \rightarrow \infty.
$$
\begin{figure
\centering
\includegraphics[width=0.8\textwidth]{static_new.eps}
\caption{Soliton interaction with the defect with small strength $\left(\epsilon = 0.1\right)$, where initial velocity of the soliton is zero. The soliton is trapped and an oscillatory movement is observed.}
\label{fig:soliton0}
\end{figure}
In this paper we consider $V$ as the stochastic variable and we choose a collocation nodal set $\left\lbrace V^{j}, \alpha ^{j}\right\rbrace_{j=0}^{Q}$ in space $\Gamma$, where $V^{j}$ are the $j$th collocation points and $\alpha^{j}$ the corresponding weights.
For each $j = 0,\cdots, Q$, we solve the problem given by Eqs. \ref{GP1} and \ref{initial1} with the parameters $\epsilon$ and $V^{j}$ and let the solution set be $\left\lbrace u_{0},\cdots , u_{Q}\right\rbrace$ where $u_{j}$ is the solution for $V = V_{j}$. For solving this deterministic equation, we employ the high-order SSFM. The approximate gPC expansion coefficients are
$$
\hat{u}_{m}(x,t) = \sum_{j=0}^{Q} u_{j}\left(x,t,V_{j}\right)\phi_{m}\left(V_{j}\right)\alpha_{j}, \; m = 0,\cdots , Q,
$$
where $\left\lbrace \phi_{m}\right\rbrace$ are the orthonormal polynomials. And finally we construct the $Q$th order gPC approximation
$$
u(x,t; V) \approx \sum_{m=0}^{Q}\hat{u}_{m}\left(x,t\right)\phi_{m}\left(V\right), \; \mathrm{where}\; V = \left\lbrace V_{0}, V_{2}, \cdots , V_{Q}\right\rbrace.
$$
\section{gPC collocation algorithm for solving NLSE}
\begin{figure
\centering
\includegraphics[width=0.6\textwidth]{unperturbate_nls_soliton.eps}
\includegraphics[width=0.6\textwidth]{128.eps}
\includegraphics[width=0.60\textwidth]{V003eps_05.eps}
\caption{Top: Soliton solution without any defect. Middle: Soliton passes through the defect with the initial velocity $V = 0.001$ and the defect amplitude $\epsilon = 0.1$. Bottom: Soliton trapped by the defect with the initial velocity $V = 0.003$ and the defect amplitude $\epsilon = 0.5$}
\label{fig:soliton1}
\end{figure}
The following algorithm describes how to calculate the critical velocity by using the gPC collocation method.
We use the gPC method to find the critical velocity $V_{c}$ efficiently for any given $\epsilon$. Here the soliton velocity $V$ is the stochastic variable. Suppose we know in advance that the critical velocity $V_{c}$ lies between $V_{a}$ and $V_{b} \; \left(V_{a} < V_{b}\right)$ and consider $V$ has a uniform distribution over $\left[V_{a}, \; V_{b}\right]$. Since the distribution is uniform, we use the Legendre polynomials for expanding the solution in the random space. For this purpose we choose $N+1$ Gauss-Legendre quadrature points with the weights. Let the set $\left\lbrace \alpha_{i},\;\omega_{i}\right\rbrace_{i=0}^{N}$ describe the $\left(N+1\right)$ quadrature points $\alpha_{i}$ and the corresponding weights $\omega_{i}$.
Now find the solution of Eq. \ref{GP1} for each $V = \alpha_{i}$ by using the high-order SSFM. For this purpose one must use a sufficiently large computational domain and sufficiently long time interval. We set up the domain size and the computational time in such a way that no solution leaves the domain yet with the given final time. For example when $\epsilon = 0.3$, we use the domain size $\left[-L, \;L\right]= \left[-40, \; 40 \right]$ and the final time $t_{f} = 12000$. We are solving the NLSE for $u_{j}\left(x, t, V_{j}\right)$ for all $V_{j}$ with the same final time.
We reconstruct the soliton solution for each simulation for $ x \in \left[-L, \; L^{'}\right]$ at the final time. $L^{'}$ is chosen in such a way that only the trapped solutions exist inside $\left[-L, \; L^{'}\right]$. We know if the solution is trapped, it would stay around the position of the defect (in our case at $x = 0$). So $L^{'}$ must be close to zero. In our computation, we choose $L^{'}$ where the mean solution vanishes near $x = 0^{+}$. For the $i$-th quadrature point $\alpha_{i}$, we denote the solution by $u_{i}\left(x,T_{f}, \alpha_{i}\right)$.
\begin{figure
\centering
\includegraphics[width=0.47\textwidth]{pass_epsilon1_v00009.eps}
\includegraphics[width=0.47\textwidth]{pass_epsilon08_v00008.eps}
\includegraphics[width=0.47\textwidth]{eps_45_trapped.eps}
\includegraphics[width=0.47\textwidth]{eps_45_passed.eps}
\caption{Top left: Soliton transmitted through the defect when $\epsilon = 0.08$ and $V = 5 \times 10^{-5}$. Top right: Soliton transmitted through the defect when $\epsilon = 0.1$ and $V = 8 \times 10^{-5}$. Bottom left: Soliton is trapped by the defect when $\epsilon = 4.5$ and $V = 0.220048$.
Bottom right: Soliton is transmitted through the defect when $\epsilon = 4.5$ and $V = 0.23995187$. For $\epsilon = 4.5$, radiation effect is clearly visible. }
\label{fig:soliton3}
\end{figure}
\begin{figure
\centering
\includegraphics[width=0.60\textwidth]{poincare.eps}
\includegraphics[width=0.60\textwidth]{poincare2.eps}
\includegraphics[width=0.60\textwidth]{slow_vel_high_eps.eps}
\caption{The nonlinear interaction of the soliton with the defect (the dotted line). Top: The nonlinear interaction is prominent when the soliton hits the deffect $\left(\epsilon = 0.3\right)$ with small velocity $\left(\sim 10^{-5}\right)$ compared to the interaction with the high velocity $\left(\sim 10^{-1}\right)$ where $\epsilon = 4.5$ (middle). Bottom: The interaction of slowly moving soliton $\left(V \sim 10^{-5}\right)$) with the defect with high value of $\epsilon \left( = 4.5\right)$.}
\label{fig:poincare}
\end{figure}
Evaluate the approximate gPC expansion coefficients by
$$
\hat{u}_{m}\left(x, T_{f}\right) = \sum_{i=1}^{Q}u_{i}\left(x,T_{f},\alpha_{i}\right) L_{m}\left(\alpha_{i}\right)\omega_{i},
$$
where $\left\lbrace L_{m}\right\rbrace_{m = 0}^{Q}$ is the set of Legendre polynomials and $\omega_i$ are the quadrature weights. The full gPC solution is given by
\begin{equation}
u(x,T_{f},\alpha) = \sum_{k=0}^{Q}\hat{u}_{k}\left(x,T_{f}\right)L_{k}(\alpha).
\label{sol_gpc1}
\end{equation}
The mean solution is given by the 1st mode \cite{Xiu2}, i.e
\begin{equation}
\hat{u}_{0}\left(x, T_{f}\right) = \sum_{i=0}^{Q}u_{i}\left(x,T_{f},\alpha_{i}\right)L_{0}\left(\alpha_{i}\right)\omega_{i}.
\label{sol_gpc_mean}
\end{equation}
From Eq. \ref{sol_gpc_mean}, one can construct the average energy $\bar{E}$ of the system between $\left[-L, \; L^{'}\right]$ and $\left[-L, \; L\right]$ at the final time, that is,
\begin{eqnarray}
\bar{E}_{L} =\frac{1}{2} \int_{-L}^{L}\left|\hat{u}_{0}\left(x, T_{f}\right)\right|^{2}dx,\; \bar{E}_{L^{'}} =\frac{1}{2} \int_{-L}^{L^{'}}\left|\hat{u}_{0}\left(x, T_{f}\right)\right|^{2}dx.
\label{definition}
\end{eqnarray}
Suppose that among $N$ solutions, $N_{1}$ solutions are trapped inside $\left[-L, \;L^{'} \right]$.
Then $N_{1}$ can be estimated for large $N \rightarrow \infty$ by
\begin{eqnarray}
\frac{N_{1}}{N} = \frac{\bar{E}_{L^{'}}}{\bar{E}_{L}}, \quad \frac{N_{1}}{N} = \frac{V_{c}-V_{a}}{V_{b}-V_{a}}, \nonumber
\end{eqnarray}
where $V_{c}$ is the critical velocity for given $\epsilon.$ So $V_{c}$ is evaluated by
\begin{equation}
V_{c}= V_{a} + \left(V_{b} - V_{a}\right) \frac{\bar{E}_{L^{'}}}{\bar{E}_{L}}.
\label{v_critical}
\end{equation}
If we increase the number of quadrature points, then the critical velocity can be determined more accurately. For our simulations we used $24$ Gauss Legendre quadrature points and obtained spectral accuracy of $\sim 10^{-12}$.
Figure \ref{fig:soliton6} shows the spectral convergence of the error of the critical velocities with the increasing number of the quadrature points.
\noindent
\textbf{Remark:}
\noindent
\textit{The solution $u(x,t,V)$ has possibly a jump at $V = V_c$ for $t \rightarrow \infty$ because of the critical behavior of the soliton solution around the potential. This means that the spectral reconstruction of $u(x,t,V)$ for any $V \in [V_a, V_b]$ using ${\hat u}_l (x,t), l = 0, \cdots, Q$ may fail to converge to the right solution due to the discontinuity at $V = V_c$. This was also addressed in our previous work for the critical behavior of the soliton solution for the sine-Gordon equation \cite{Chakraborty_Jung}. Here note that the proposed method in this paper uses only the first moment ${\hat u}_0(x,t)$ to estimate the critical velocity $V_c$ but not the reconstruction of $u(x,t,V)$. The mean solution, ${\hat u}_0(x,t)$ is convergent. }
In Eq. \ref{v_critical}, the convergence of $V_c$ mainly depends on $R := \frac{\bar{E}_{L^{'}}}{\bar{E}_{L}}$. As the definition in Eq. \ref{definition}, the convergence of $R$ then depends on how ${\hat u}_0(x,t)$ converges with $N$. In our previous work \cite{JungSong}, it was proven that ${\hat u}_0(x)$ converges fast enough although the original function $u(x,V)$ is discontinuous in the random variable $V$. As we will discuss in the next section, numerical results in Section 5 (Figure 7) implies that $R$ shows spectral convergence with $N$.
\begin{figure
\centering
\includegraphics[width=0.47\textwidth]{avg_eps_03.eps}
\includegraphics[width=0.47\textwidth]{avg_eps_03_zoom.eps}
\includegraphics[width=0.47\textwidth]{avg_sol_eps27.eps}
\includegraphics[width=0.47\textwidth]{avg_sol_eps27_zoom.eps}
\includegraphics[width=0.55\textwidth]{hermite_avg_eps03.eps}
\caption{ First mode (mean) of the gPC expansion for different $\epsilon$. Top and middle: The Legendre chaos. Bottom: The Hermite chaos. Right figures of the top and middle panels show the locations of $L^{'}$ for different $\epsilon$. }
\label{fig:soliton4}
\end{figure}
\section{Numerical results}
\begin{figure
\centering
\includegraphics[width=0.8\textwidth]{v_eps_plot.eps}
\caption{ Critical Velocity vs. $\epsilon$ where $\epsilon \in \left[0.05, \; 4.5\right]$. }
\label{fig:soliton5}
\end{figure}
We first consider the high value of $\epsilon$, say $\epsilon = 2.7$. By doing few Monte-Carlo simulations we roughly estimate the interval $V \in \left[V_{a},\; V_{b}\right], \; V_{c} \in \left[V_{a},\; V_{b}\right] $ where $V_{c}$, the critical velocity may be located. For $\epsilon = 2.7$, we use $V_{a} = 0.1$ and $V_{b} = 0.14$. Since for moderate and high values of $\epsilon$, the simulation time is relatively less than the simulation time with smaller range of $\epsilon$, we follow the same procedure to find the suitable intervals. But for the small value of $\epsilon$, i.e. $\epsilon < 1.0$, where the simulation time is long, we use the extrapolation of $V_{c}$ from the previous $\epsilon$ to get the rough estimate of the interval.
To apply the gPC collocation method, one also needs to find the value of $L^{'}$. We do not have any fixed $L^{'}$ which can serve for all $\epsilon$. Instead, we have different $L^{'}$ for different $\epsilon$. A heuristic approach is used to find $L^{'}$. For the given value of $\epsilon$, we construct the mean solution by Eq. \ref{sol_gpc_mean}. Since some solutions are trapped and some of them are transmitted, there are few bumps near the defect and few bumps are far from the defect. Clearly there exists a separation point between these two groups of bumps. Ideally the $x-$coordinate of this point would be zero but due to the domain truncation, radiation effect etc. it may not be equal to zero. By observing the graph carefully we can easily find the separation point which we use as $L^{'}$. For the small and moderate values of $\epsilon$, determining accurately $L^{'}$ is easy, but for the high values of $\epsilon$, we need extra care. For the high value of $\epsilon$, the values of $V_{a}$, $V_{b}$ are also high and we can not run the simulations for a long time because some solutions may leave the domain and re-enter the domain from the other side due to the periodic boundary conditions. So in this case we need to study the bumps carefully to locate $L^{'}$. In Figure \ref{fig:soliton4}, the zoomed graphs of the mean solution of each $\epsilon$ are given in the right panel of the top and middle figures. We find that for $\epsilon = 0.3, \; L^{'} = 12$ and for $\epsilon = 0.5, \; L^{'} = 13.5$. Similarly for $\epsilon = 2.7, \; L^{'} = 10$ and for $\epsilon = 3.0, \; L^{'} = 15$.
Figure \ref{fig:soliton0} presents the interaction of the soliton with the $\delta$-function. Here we choose the initial velocity, $V_{0} = 0$ and the potential strength $\epsilon = 0.1$. The soliton is located at $x_{0} = -0.3$ initially, which is inside the influence zone of the potential. The nonlinear interaction is observed and the soliton solution exhibits an oscillatory behavior along the line $x = 0$. This case was discussed in \cite{Holmer1, Holmer2}. But such an initial condition may not necessarily satisfy the given equation. The initial position of the soliton must be out of the influence zone of the potential and the soliton must be allowed to move freely before it hits the defect.
In all cases we consider the starting point of the soliton $\left(x_{0}\right)$ is far from the position of $\delta$-function, i.e. outside the influence region of the potential.
Figure \ref{fig:soliton1} shows the behavior of the soliton solutions in three different cases. When $\epsilon = 0$, that is the case when there is no $\delta$-function, the soliton solution passes unperturbedly. But for nonzero $\epsilon$, the soliton behaviour depends on its initial velocity. For $\epsilon = 0.1$, the soliton passes through the defect for $V = 0.001$ and for $\epsilon = 0.5$ and $V = 0.003$, soliton is trapped by the defect. For both cases, the soliton passed or trapped as a whole. There is no radiation due to the small soliton velocities \cite{Cao_Malomed}.
Figure \ref{fig:soliton3} represents the long time simulations for $\left(\epsilon,\;V\right) = \left(0.08,\; 5\times 10^{-5}\right)$ (top left), $\left(0.1,\; 8\times 10^{-5}\right)$ (top right), $\left(4.5,\; 0.220048\right)$ (bottom left) and $\left(4.5,\; 0.23995187\right)$ (bottom right).
For the case that $\epsilon$ is small and $V$ is also very small accordingly, the soliton is transmitted through the defect without any radiation. But for the high value of $\epsilon$, usually greater than $2.7$, where the critical velocity is also high, the radiation effect is observed due to the soliton-defect interaction. The bottom panel of Figure \ref{fig:soliton3} exhibits the radiation effect for $\epsilon = 4.5$. For both the ``trapped" and ``transmitted" situations, the radiation effect is observed. The bound state effect is also observed in the bottom right, the details of which was discussed in \cite{Holmer}.
Figure \ref{fig:poincare} shows the nonlinear interactions of the soliton with different soliton velocities. When the soliton velocity is small, nonlinear property dominates as shown in Figure \ref{fig:soliton3}. During the time of interaction with the defect (the dotted line), the soliton velocity increases and after crossing the defect, the velocity turns into its previous value. When the soliton velocity is high, the linear effect dominates and the soliton velocity does not changes during the collision but the direction of the propagation changes. That is the soliton continues its motion with the same velocity.
When a slowly moving soliton hits the defect with high strength $\left(\epsilon = 4.5\right)$, the soliton is trapped by the defect but due to the nonlinear interactions, radiations and transmissions are also seen (bottom figure).
Figure \ref{fig:soliton4} shows the mean solutions at the final time. This is the first mode of the solution by the gPC collocation method. Here we used $V$ as a stochastic variable, $V \in \left[V_{a}, \; V_{b} \right]$ and $V_{a}$ and $V_{b}$ are different for different values of $\epsilon$. We used both the Legendre and Hermite chaos. We need to consider the uniform distribution and normal distribution for the Legendre and Hermite chaos respectively. In Figure \ref{fig:soliton4}, the figures in the top panel are obtained using the Legendre chaos for $\epsilon = 0.3,\; 0.5$.
Those solitons that are trapped by the defect are confined around the position of the defect. In our case, the defect, the $\delta$-function is located at $x = 0$. There are multiple peaks in the mean solution, but around $x = 0$ the peaks are higher than the others, which implies that some solitons are trapped, and the rest are transmitted. These figures are used to locate the position of $L^{'}$. If we see the zoomed figure in the right panel, we easily locate $L^{'}$ for different $\epsilon$.
For the middle panel figures in Figure \ref{fig:soliton4}, we plotted the mean solutions and zoomed one for $\epsilon = 2.7,\; 3.0 \; \mathrm{and}\; 4.5$. The sharp peaks at $x = 0$ imply that the most of the solutions are trapped in that range of $V$ and some of them are transmitted. We already mentioned that in this region of such a large value of $\epsilon$, the radiation effects are visible, which are also showed in the figure. The values of $L^{'}$ are pointed for different $\epsilon$ values in the figure.
Same explanation for $\epsilon = 0.5$.
Next we consider the case that $V$ is normally distributed and we use Hermite polynomials \cite{XiuBook} and the Gauss-Hermite quadrature points \cite{Gottlieb}. Let $V_{a} = \alpha, \; V_{b} = \beta$ and $V \in \left[\alpha, \; \beta \right] $, $\xi \in \left[-1, \; 1 \right]$, $\gamma \in \left(-\infty, \; \infty \right) $. The linear transformation between $V$ and $\xi$ is given by
$$ V(\xi) = \left(\frac{\beta - \alpha}{2} \right)\xi + \frac{1}{2}(\alpha + \beta) $$
and the transformation between $\xi$ and $\gamma$ is given by \cite{Chen_Gottlieb}
\begin{eqnarray}
\gamma &=& \frac{\xi}{1-\xi^2}, \qquad \xi \ne 0, \nonumber \\
&=& 0, \qquad \qquad \xi = 0. \nonumber
\end{eqnarray}
Or we have,
\begin{eqnarray}
\xi &=& \frac{-1+\sqrt{1+4\gamma^2}}{2\gamma}, \qquad \gamma \ne 0, \nonumber \\
&=& 0, \qquad \qquad \qquad \gamma = 0. \nonumber
\end{eqnarray}
Thus we have,
\begin{equation}
V(\gamma) = \left(\frac{\beta - \alpha}{2} \right)\left[\frac{-1+ \sqrt{1 + 4\gamma^2}}{2\gamma} \right]+\frac{1}{2}(\alpha + \beta), \nonumber
\label{trans1}
\end{equation}
where $\gamma $ has the normal distribution with mean $0$ and the standard deviation (SD) $0.1$.
For the simulation we consider $\epsilon = 0.3$ and $V \sim N\left[0,\;0.1\right]$. The figure in the bottom panel of Figure \ref{fig:soliton4} shows the mean solution at the final time obtained by the Hermite chaos. Although the mean solutions obtained from the Legendre and Hermite chaos are different, we observe that the location of $L^{'}$ is same for both cases.
Using a series of those simulations above for different values of $\epsilon$ where $\epsilon \in \left[0.05, \; 4.5\right]$, we determine the critical velocities with respect to different $\epsilon$.
The results are plotted in semi-logarithmic scale in Figure \ref{fig:soliton5}. It is observed that for the small values of $\epsilon$ where $\epsilon < 0.1$, the curve is very stiff and the slope changes sharply around $\epsilon =0.1$. From $\epsilon > 0.1$, the curve increases steadily.
The \textit{``trapped"} and the \textit{``untrapped"} regions are clearly shown in the figure. The $V - \epsilon$ graph is the boundary of those two regions.
{{
\begin{table}
\label{table1}
\caption{Convergence of $V_{c}$ with $N$ for the Legendre Chaos. $\epsilon = 0.3,\;1.0$ and $4.5$.}
\begin{center}
\begin{tabular}
{cccc}\hline\em $N$ &\em & $V_{c} \times 10^{3}$ & \\\hline & $\epsilon = 0.3 $ & $\epsilon = 1.0$ & $\epsilon = 4.5$\\\hline\hline $2$ & $1.658675134594813$ &$23.83399810435849$&$233.3998104358486$ \\\hline $4$ & $1.786459011090171$ &$24.00997282851472$& $236.0392987896415$ \\\hline $8$ & $1.788091882999389$ &$24.01306953886891$& $236.7359886666223$ \\\hline $12$ & $1.788112748469627$ & $24.01312403314395$ &$236.9198791482719$ \\\hline $16$ & $1.788113015096694$ &$24.01312499210545$& $236.9684168267407$ \\\hline $20$ & $1.788113018503758$&$24.01312500898075$& $ 236.9812282904446$\\\hline $24$ & $1.788113018547295$ &$24.01312500927771$ & $236.9846098613687$ \\\hline
\end{tabular}
\end{center}
\end{table}
}}
\subsection{Convergence analysis}
We define the error of the critical velocities by
$$
\mathrm{Error}^{\epsilon}(N) = \left|V_{c}^{\epsilon}(N) - V_{c}^{\epsilon}(N-1)\right|,
$$
where $N$ is the number of collocation points. Figure \ref{fig:soliton6} shows the convergence of errors obtained by the Legendre and Hermite chaos. We do the convergence analysis for various values of $\epsilon$. We choose $\epsilon = 0.3$ (small) , $\epsilon = 1.0$ (moderate) and $\epsilon = 4.5$ (high).
For the Legendre chaos, the critical velocities for different $N$ are presented in Table $1$. For $\epsilon = 0.3$ and $\epsilon = 1.0$, we calculate the errors for both the Legendre and Hermite chaos and for $\epsilon = 4.5$ we use the Legendre chaos. For Hermite chaos, we expect to have the similar results. The graphs are plotted in semi-logarithmic scale. Figure \ref{fig:soliton6} shows all the graphs are a straight line, which confirms spectral convergence but the convergence rates are different for different cases. For $\epsilon = 0.3$ and $\epsilon = 1.0$, Hermite chaos exhibits slower convergence rate than the Legendre chaos. Also if we compare the graphs for the Legendre chaos for different cases, it is found that the convergence rate decreases with the increases of the value of $\epsilon$. That is, the smaller is the value of $\epsilon$, the faster convergence is obtained. One of the possible reasons is because of the radiation effect. As $\epsilon$ increases, the radiation effect becomes visible and it makes difficult to locate the position of $L^{'}$ accurately. According to our numerical results, our main result is stated by the following:
\textit{The numerical scheme stated in Section $4$ to find the critical velocity $V_{c}$ has the spectral convergence and the rate of convergence decreases with increase of the value of $\epsilon$.}
\begin{figure
\centering
\includegraphics[width=0.47\textwidth]{leg_error_conv.eps}
\includegraphics[width=0.47\textwidth]{leg_error_conv_eps1.eps}
\includegraphics[width=0.47\textwidth]{leg_error_conv_eps45.eps}
\caption{Spectral convergence of the critical velocities for $\epsilon = 0.3, 0.1$ and $4.5$. Graph shows the spectral convergence for both the Legendre and Hermite chaos. Note that the Legendre chaos shows faster convergence than Hermite chaos.}
\label{fig:soliton6}
\end{figure}
\section{Conclusion}
In this paper we studied the NLSE with the singular potential.
We proposed an efficient method of determining the critical soliton velocities, $V_{c}$, by using the gPC collocation method. We studied the wide range of $\epsilon$, i.e. $\epsilon \in \left[0.05,\; 4.5\right]$. For $\epsilon < 0.05$ the numerical simulations demand a huge computational time due to the very small soliton velocity $\left(V \sim 10^{-10}\right)$.
We studied the convergence analysis to prove the merit of our proposed numerical scheme. We found the spectral convergence in all cases. The main development of this paper is the use of the gPC collocation method to determine the critical velocity of the soliton for given $\epsilon$ with the desired level of accuracy. We obtained $V_c$ accurately with a small number of simulations. In our future work, we will further study the case that $\epsilon \ll 0.05$. Also for the high values of $\epsilon$, where radiation effect is prominent and the convergence of the proposed method becomes slower due to the radiation effect, an efficient numerical method dealing with this effect will be investigated.
\vskip .1in
{\bf Acknowledgement:}
\noindent
The first author is grateful to Gino Biondini for developing and implementing high-order SSFM.
|
\section{High-harmonic generation by continuum wave packets}
\label{sec:ccHHG}
High-order harmonic generation (HHG) is a key process in ultrafast science and well-understood within the three-step model~\cite{CORKUM1993}: the bound wave function of an atom is partially freed by a strong laser field, accelerated in it, and driven back to its parent ion. At that point, the ionized and bound portions of the electronic wave packet interfere, giving rise to a strong, coherent high-frequency dipole response that can lead to the emission of a HHG photon along with the recombination of the electron into the bound state.
In our work \cite{KOHLER2010}, we advance the interference model of HHG~\cite{PUKHOV2003,ITATANI2004} to provide a comprehensive physical picture including continuum--continuum (CC) transitions: any two wave packets of the same electronic wave function that have been split and have acquired different energies lead to coherent HHG emission when they simultaneously reencounter the core region. The emitted photon energy is exactly the energy difference of the two wave packets. Starting from this point of view, we found that a new CC transition plays a significant role in the over-the-barrier (OBI) ionization regime. This transition occurs when two wave packets ionized in different half cycles of a laser pulse recollide at the same time. Note that this transition is different to the CC harmonics described in~\cite{CCHarmonics}.
As a realistic model, we study the hydrogen atom in a laser field in the OBI regime and solve the three-dimensional time-dependent Schr\"odinger equation numerically~\cite{BAUER2006}. The laser pulse is shown in Fig.~\ref{fig-HHGnumerical}a and chosen such that almost complete depletion of the ground state occurs on the leading edge of the pulse.
\begin{figure}[h]
\centering
\includegraphics[width=7.0cm]{mkohler_fig2_pg}
\caption{\label{fig-HHGnumerical}
Time-frequency analysis of HHG showing the signature of CC wave-packet interference. a) Laser pulse used for the calculation (solid line, left axis) and ground-state population (dashed line, right axis). b)~Windowed Fourier transform of the HHG emission. The dashed black lines are the classically calculated kinetic energies of electrons returning to the ion and the solid red line their difference energy. Reproduced from Ref.~\cite{KOHLER2010}. Copyright (2010) by the American Physical Society.}
\end{figure}
To analyze the time-resolved frequency response of HHG, we calculate the windowed Fourier transform of the dipole acceleration obtained from the TDSE calculation and display it in Fig.~\ref{fig-HHGnumerical}b. For comparison, the two dashed black lines in the figure are the classical recollision energies for trajectories starting from two different laser half cycles, respectively, which are in agreement with the traditional continuum--bound (CB) signal whereas the red line denotes their difference.
The CC transition is evidenced by the excellent agreement of the quantum-mechanical response with the red line. Interestingly, the CC component of the dipole response is the dominant contribution for several half cycles after $t=150$~a.u. (atomic units are employed throughout). This can be understood from the fact that depletion of the ground state occurs around that time. Then, coherent HHG can only occur by the presence of the various parts of the wave function in the continuum.
Moreover, we developed a strong-field approximation model for CC HHG suitable for the OBI regime and based on the evaluation of the dipole acceleration
$\mathbf{a}(t)=-\langle \Psi ,t\vert \mathbf{\nabla} V\vert \Psi ,t\rangle$~\cite{Gordon2005} rather than the dipole moment to include the distortion of the recolliding waves by the Coulomb potential required for momentum conservation. The saddle-point approximation is applied to the expression making a computationally fast evaluation of the process possible.
This analytical model also allows to extract the HHG emission phase $\phi=S(\mathbf{p}'',t,t'')-S(\mathbf{p}',t,t')+I_p(t'-t'')-\omega_{\mathrm{H}} t$ with the classical action $S$~\cite{lewenstein}, canonical momentum~$\mathbf{p}$, HHG emission time $t$, ionization times $t'$ and $t''$, HHG frequency $\omega\In{H}$ and ionization energy $I\In{p}$.
The expression can be converted to $\phi(I)=\alpha\In{cc}I$ being a function of the laser pulse peak intensity $I$, which reveals a striking difference for CB as compared to CC transitions: the sign of $\alpha\I{CC}$ and $\alpha\In{CB}$ differs for both types of transitions allowing to separate the CC from the CB harmonics via phase-matching.
This way, the measurement of CC spectra could by employed for qualitatively advancing tomographic molecular imaging~\cite{ITATANI2004}: instead of probing the orbital shape of the active electron, the effective atomic or molecular potential mediating the transition between the two wave packets could be assessed.
\section{Emergence of a high-energy plateau in HHG from resonantly excited atoms}
\label{sec:rabiHHG}
In the last two decades, HHG in the non-relativistic regime has been developed to a reliable source of coherent extreme ultraviolet (XUV) radiation. Its advancement into the hard x-ray domain would allow for a much wider range of applications. The straightforward approach to employ larger laser intensities is demanding due to the relativistic electron drift~\cite{drift} and the large electron background that is generated by the strong laser field causing phase-mismatch~\cite{Popmintcheva:PM-09}. On the other hand, the large scale x-ray free electron lasers (XFEL) routinely generate several keVs of photon energy but are limited in coherence and, thus, sub-femtosecond pulses with this technique are not in reach at the moment.
We show~\cite{buth:ra-11} that by combining both, the HHG process and radiation from an XFEL, coherent light pulses can be obtained having the extremely short time structure of the HHG and photon energies larger than the XFEL.
In addition to the increase of the HHG photon energy, the scheme can be employed for ultrafast time-dependent imaging~\cite{morishita2008} involving inner shells and for the characterization of the x-ray pulse of the XFEL.
The proposed scheme works as follows: atoms are irradiated by both an intense optical laser field and an x-ray field from a FEL. The x-ray energy is chosen to be resonant with the transition between the valence and a core level in the cation. As soon as the valence electron is tunnel ionized by the optical laser field, the core electron can be excited to the valence vacancy. Then the continuum electron, returning after a typical time of 1 fs, can recombine with a core hole rather than with the valence hole from that it was previously tunnel ionized and thus emit a much higher energy.
We developed an analytical formalism to cope with the two-electron two-color problem~\cite{buth:ra-11,buth:pra-11}. A two-electron Hamiltonian is constructed mostly from tensorial products of one-electron Hamiltonians that describe the
losses due to tunnel ionization and direct x-ray ionization via phenomenological decay constants in conjunction with Auger decay of the intermediate hole.
The system is described by equations of motions based on the Schr\"odinger equation and the solutions can be found in the dressed state basis.
We apply our theory to the $3d \rightarrow 4p$ resonance in a krypton cation as well as to the $1s \rightarrow 2p$ resonance in a neon cation. The results for a resonant sinusoidal x-ray field for two different intensities are shown in Fig.~\ref{fig:Rabi-cont}. The chosen optical laser field intensity is $3\times10^{14}\U{W/cm^2}$ for krypton and $5\times10^{14}\U{W/cm^2}$ for neon both at $800\U{nm}$ wavelength.
\begin{figure}[h]
\begin{center}
\includegraphics[width=0.49\textwidth]{rabi_Kr_fig3a}\hskip 0.2cm
\includegraphics[width=0.49\textwidth]{rabi_Ne_fig3b}\\
\includegraphics[width=0.49\textwidth]{rabi_Kr_fig3c}\hskip 0.2cm
\includegraphics[width=0.49\textwidth]{rabi_Ne_fig3d}
\caption{HHG photon numbers of the $h^{\textrm{th}}$ harmonic for different x-ray intensities from (a), (c) krypton and (b), (d) neon. The solid black line stands for the recombination to the valence state whereas the more energetic plateau arising from core hole recombination is red the dashed line. The thin lines in the spectrum are obtained by neglecting ground-state depletion due to direct x-ray ionization. The xuv and laser pulse durations are three optical laser cycles in all cases.} \label{fig:Rabi-cont}
\end{center}
\end{figure}
The most striking feature in the obtained spectra is the appearance of a second plateau. It is upshifted in energy with respect to the first plateau by the energy difference between the two involved core and valence states. The two plateaus have comparable harmonic yields for x-ray intensities above $10^{16}\U{W/cm^2}$. Quite importantly, the losses due to x-ray ionization do not lead to a significant drop of the HHG rate as can be seen by comparing the thick and thin lines. Both lines almost coincide for neon due to the negligible depletion induced by ionization. The second plateau bears signatures of the core state and may offer a route for ultrafast time-dependent chemical imaging of inner shells~\cite{ITATANI2004,morishita2008}. Moreover, by exploiting the upshift in energy, attosecond x-ray pulses come into reach.
\section{High-harmonic generation without attochirp}
\label{sec:attoHHG}
A particularly fascinating property of HHG is its time structure enabling for the generation of extremely short pulses~\cite{GOULIELMAKIS2008,SANSONE2010}. Nowadays pulses down to a duration of 63~as~\cite{KO2010} have been generated and the bandwidth to generate pulses of only 11~as is available~\cite{CHEN2010}. The emitted pulses have an intrinsic chirp, the so-called attochirp~\cite{MAIRESSE2003,KAZAMIAS2004} and, thus, are much longer than their bandwidth limit. To compensate the attochirp, dispersive optical media~\cite{KO2010,chirp-compensation} are employed.
An alternative way to circumvent this problem would be to modify the HHG process such that the light is emitted without attochirp.
It is shown~\cite{KOHLER:CF-11} that by means of laser pulse shaping, employing soft x rays for ionization~\cite{xuv-ass-HHG} and using an ionic gas medium, attosecond pulses with arbitrary chirp can be formed including the possibility of attochirp-free HHG and bandwidth-limited attosecond pulses.
\begin{figure
\begin{center}
\includegraphics[width=0.5\textwidth]{as_schematic_a}\\
\hskip0.0cm\includegraphics[width=0.5\textwidth]{as_schematic_b}
\caption{Schematic of the recollision scenario: a) A half cycle of the tailored laser field (black). The red line is the assisting x-ray pulse. b) Different one-dimensional classical trajectories in the field of (a) which start into the continuum at different times but revisit the ionic core at the same time. Reproduced from Ref.~\cite{KOHLER:CF-11}. Copyright (2011) by the Optical Society of America.} \label{figopti}
\end{center}
\end{figure}
The principles are illustrated in the trajectory picture of HHG~\cite{CORKUM1993,lewenstein} in Fig.~\ref{figopti}. Since the recollision time of a certain harmonic can be identified with its group delay~\cite{MAIRESSE2003,KAZAMIAS2004} in the emitted pulse, a simultaneous recollision of all trajectories would lead to a bandwidth-limited attosecond pulse. The demand of simultaneous recollision can be fulfilled if the electron is freed by single-photon ionization when the x-ray frequency $\omega\In{X}$ is much larger than the binding energy. In this case the electron has a large initial kinetic energy directly after ionization. Let us focus on the two example trajectories marked by $\alpha$ and $\beta$ in Fig.~\ref{figopti}b. Both are ionized at instants separated by a small time difference $\delta t\In{i}$. The chosen starting direction of the trajectories along the laser polarization direction is such that they are subsequently decelerated by the laser field and eventually recollide. Note that the velocity difference between $\alpha$ and $\beta$ is conserved in time for a homogeneous laser field. With a convenient choice of the parameters, the velocity difference between both acquired during $\delta t\In{i}$ can be such that both recollide simultaneously.
In the following, we exemplify our method in two cases producing bandwidth-limited pulses below 10~as and 1~as.
The optimized optical laser pulses (frequency $\omega=0.06$a.u.) are shown in Fig.~\ref{fig_highE}a and c and the parameters are indicated in Table~\ref{tab_values}.
\begin{figure}[th]
\begin{center}
\includegraphics[width=0.9\textwidth]{kohler_as-shape_fig3}
\caption{a) the laser field composed of 8 Fourier components that illuminates an Li$^{2+}$ ion with $I\In{p}=4.5$~a.u.. The parameters of the additional x-ray field are indicated in the first row of Table \ref{tab_values}. The resulting 8~as pulse is shown in b). c) displays the laser field needed to create a pulse of 800~zs duration from Be$^{3+}$ ions with $I\In{p}=8$~a.u. The parameters are indicated in the second row of Table~\ref{tab_values}. The respective pulse is shown in d). The dashed red lines are the temporal phases of the pulses. Reproduced from Ref.~\cite{KOHLER:CF-11}. Copyright (2011) by the Optical Society of America.} \label{fig_highE}
\end{center}
\end{figure}
The attosecond pulses are bandwidth limited as can be seen from the spectral phases (red lines in Fig.~\ref{fig_highE}b and d).
\begin{table}[h]
\begin{center}
\begin{tabular}{c|c|c|c|c|c}
$N_{F}$&$I_L$[W/cm$^2$]&$\omega_x$[eV]&$I_X$[W/cm$^2$]&ion&$I_p$ [a.u]\\
\hline
8&$10^{16}$&218&$3.5\times 10^{14}$&Li$^{2+}$&4.5\\
\hline
20&$10^{17}$&996&$1.4\times10^{15}$&Be$^{3+}$&8\\
\hline
\end{tabular}
\caption{Parameters for the two examples in Fig.~\ref{fig_highE}: $N\In{F}$ represents the number of Fourier components contained in the tailored driving pulse, $I\In{L}$ its peak intensity, $\omega\In{X}$ the x-ray frequency employed for ionization, $I\In{X}$ its intensity and the ionization energy $I\In{p}$.}\label{tab_values}
\end{center}
\end{table}
However, the method still has several drawbacks. It is experimentally demanding to create a pure ionic gas and to achieve phase-matching in a macroscopic medium due to the free electron dispersion. Moreover, the required large initial momentum and the dipole angular distribution of the ionization process lead to an increased spread of the ionized wave packet as compared to tunnel ionization. Further, the ionization rate is small due to $\omega\In{X}\gg I\In{p}$.
Precise shaping of intense driver fields with intense harmonics as well as the synchronization of the x-ray and IR pulses is also demanding. A potential route to overcome the former difficulties could be to derive the x-rays, the IR light and its Fourier components from the same FEL electron bunch.
\section{Resonant photoionization involving two atoms}
\label{sec:ICD}
Interatomic electron--electron correlations are responsible for a variety of interesting phenomena, ranging from dipole--dipole interactions in cold quantum gases to F\"orster resonances between biomolecules. They may also lead to very characteristic effects in the photoionization of atoms.
As an illustrative example, let us consider resonant two-photon ionization in a system consisting of two hydrogen atoms~\cite{MUELLER2011}. The atoms are assumed to be separated by a sufficiently large distance $R$ so that one may indeed speak about individual atoms. The system interacts with a monochromatic laser field whose frequency is resonant with the $1s$--$2p$ transition in hydrogen. In this situation there are two quantum pathways for ionization: (i) the direct channel where a single atom is ionized by absorbing two photons from the field, without any participation of the neighboring atom; and (ii) an interatomic channel where both atoms are first excited to the $2p$ level by absorbing one photon each and afterwards the doubly excited two-atom state decays via so-called interatomic Coulombic decay (see~\cite{ICD} for recent reviews). In other words, one of the atoms deexcites and transfers its energy radiationlessly to the second atom, this way causing its ionization.
Leading to the same final state, the direct and two-center ionization pathways show quantum mechanical interference. It becomes manifest in the photoelectron spectra which consist of four lines due to the Autler-Townes splitting of the atomic levels in the external field~\cite{MUELLER2011}. Moreover, the two-center channel can be remarkably strong and even dominate over the direct channel by orders of magnitude. In fact, the ratio $[d/(R^3E)]^2$ determines the strength of the former with respect to the latter, where $d$ is the $1s$--$2p$ transition dipole moment and $E$ denotes the laser field strength. For example, at $R=1$\,nm and $E=10^4$\,V/cm this ratio is of the order of $10^4$.
Interatomic resonant photoionization may also occur in hetero-atomic systems~\cite{MARPE,NAJJARI2010}. In this case, one of the centers is first excited by single-photon absorption and subsequently, upon deexcitation, the partner atom is ionized. Evidently, this mechanism requires an hetero atomic system with an excitation energy of the one center exceeding the ionization potential of the other.
The interplay between resonant laser fields and interatomic electron-electron correlations can thus give rise to interesting and rather unexpected effects.
\\
C.B.~and M.C.K.~were supported by a Marie Curie International Reintegration
Grant within the 7$^{\mathrm{th}}$~European Community Framework Program
(call identifier: FP7-PEOPLE-2010-RG, proposal No.~266551).
C.B.'s work was funded by the Office of Basic Energy Sciences,
Office of Science, U.S.~Department of Energy, under Contract
No.~DE-AC02-06CH11357. T.P. acknowledges support by an MPRG grant of the Max-Planck Gesellschaft.
This work was supported in part by the Extreme Matter Institute EMMI.
\input{referenc1}
\end{document}
|
\section{Introduction}\label{sec1}
A \textit{convex lattice polygonal line} $\varGamma$ is a piecewise
linear path on the plane, starting at the origin $0=(0,0)$, with
vertices on the integer lattice $\ZZ^2_+:=\{(i,j)\in\ZZ^2:
i,j\ge0\}$, and such that the inclination of its consecutive edges
strictly increases staying between $0$ and $\pi/2$. Let $\CP$ be the
set of all convex lattice polygonal lines with finitely many edges,
and denote by $\CP_{n}\subset\CP$ the subset of polygonal lines
$\varGamma\in\CP$ whose right endpoint $\xi=\xi_\varGamma$ is fixed
at $n=(n_1,n_2)\in\ZZ^2_+\mypp$.
The \emph{limit shape}, with respect to a probability measure
$\PP_n$ on $\CP_n$ as $n\to\infty$, is understood as a planar curve
$\gamma^*$ such that,
for any $\varepsilon>0$,
\begin{equation}\label{eq:LLN}
\lim_{n\to\infty}\PP_n\{\varGamma\in\CP_n:
\,d\myp(\tilde{\varGamma}_n,\gamma^*)\le\varepsilon\}=1,
\end{equation}
where $\tilde{\varGamma}_n=S_n(\varGamma)$, with a suitable scaling
$S_n:\RR^2\to\RR^2$, and $d(\cdot,\cdot)$ is some metric on the path
space, e.g., induced by the Hausdorff distance between compact sets,
\begin{equation}\label{eq:dH}
d_{\mathcal H}(A,B):=\max\left\{\max_{x\in A}\min_{y\in B}|x-y|,
\,\max_{y\in B}\min_{x\in A}|x-y|\right\}.
\end{equation}
\begin{remark}
By definition, for a polygonal line $\varGamma\in\CP_n$ the vector
sum of its consecutive edges equals $n=(n_1,n_2)$; due to the
convexity property, the order of parts in the sum is uniquely
determined. Hence, any such $\varGamma$ represents a \textit{strict
vector partition} of $n\in\ZZ_+^2$ (i.e., without proportional
parts; see \cite{V1}). For ordinary one-dimensional partitions, the
limit shape problem is set out for the associated Young diagrams
\cite{V3,Yakubovich}.
\end{remark}
Of course, the limit shape and its very existence may depend on the
probability law $\PP_n$. With respect to the uniform distribution on
$\CP_n$, the problem was solved independently by Vershik \cite{V1},
B\'ar\'any \cite{Barany} and Sinai \cite{Sinai}, who showed that,
under the scaling $S_n:(x_1,x_2)\mapsto (x_1/n_1,x_2/n_2)$ and with
respect to the Hausdorff metric $d_\calH$, limit (\ref{eq:LLN})
holds with $\gamma^*$ given by a parabola arc defined by the
equation
\begin{equation}\label{eq:gamma0}
\sqrt{1-x_1}+\sqrt{x_2}=1,\qquad 0\le x_1,x_2\le1.
\end{equation}
Recently, Bogachev and Zarbaliev \cite{BZ4,BZ-DAN} proved that the
same limit shape $\gamma^*$ appears for a large class of measures
$\PP_n$ of the form
\begin{equation}\label{eq:P-r}
\PP_n(\varGamma):=\frac{b(\varGamma)}{\displaystyle B_n}
\myp,\qquad\varGamma\in\CP_n\myp,
\end{equation}
with
\begin{equation}\label{eq:b-Gamma}
b(\varGamma):=\prod_{e_i\in\varGamma} b_{\ell_i},\qquad B_n:={}\sum_{\varGamma\in\CP_n}
b(\varGamma),
\end{equation}
where the product is taken over all edges $e_i$ of
$\varGamma\in\CP_n$, $\ell_i$ is the number of lattice points on the
edge $e_i$ except its left endpoint, and
\begin{equation}\label{eq:b-k-r}
b_\ell:=\binom{r+\ell-1}{\ell}=\frac{r(r+1)\cdots(r+\ell-1)}{\ell!}\myp,\qquad
\ell=0,1,2,\dots.
\end{equation}
This result has provided first evidence in support of a conjecture
on the limit shape universality, put forward independently by
Vershik \cite[p.\;20]{V1} and Prokhorov \cite{Prokhorov}.
The goal of the present paper is to show that the limit shape
$\gamma^*$ given by (\ref{eq:gamma0}) is universal in a much wider
class of probability measures of the form (\ref{eq:P-r}). For
instance, along with the uniform measure on $\CP_n$ this class
contains the uniform measure on the subset $\check{\CP}_n\subset
\CP_n$ of polygonal lines that do not
have
any integer points other than vertices. More generally, measures
covered by our method include (but are not limited to) direct
analogues of the three classical meta-types of decomposable
combinatorial structures
--- multisets, selections and assemblies
\cite{ABT,AT,GSE}
(see examples in Section \ref{sec2.3} below). Let us stress,
however, that our universality result is in sharp contrast with the
one-dimensional case, where the limit shape of Young diagrams
associated with integer partitions heavily depends on the
distributional type (see \cite{Bogachev,EG,V3,Yakubovich}). This
suggests that the limit shape of strict vector partitions is a
relatively ``soft'' property as compared to a more demanding case of
(one-dimensional)
integer partitions.
Let us state our result more precisely. Using the tangential
parameterization of convex paths (see \cite[\S\myp{}A.1]{BZ4}), let
$\tilde\xi_n(t)$ denote the right endpoint of part of the scaled
polygonal line $\tilde\varGamma_n=S_n(\varGamma)$ where the tangent
slope (wherever it exists) does not exceed $t\in[0,\infty]$.
Similarly, a tangential parameterization of the parabola arc
$\gamma^*$ (see (\ref{eq:gamma0})) is given by
\begin{equation}\label{eq:g*}
g^*(t)=\left(\frac{t^2+2t}{(1+t)^2}\myp, \frac{t^2}{(1+t)^2}\right),
\qquad 0\le t\le\infty.
\end{equation}
The \textit{tangential distance} between $\tilde\varGamma_n$ and
$\gamma^*$ is defined as
\begin{equation}\label{eq:d-T}
d_{\mathcal T}(\tilde\varGamma_n,\gamma^*):= \sup_{0\le
t\le\infty}\bigl|\myp\tilde\xi_n(t)-g^*(t)\bigr|.
\end{equation}
It is known \cite[\S\myp{}A.1]{BZ4} that the Hausdorff distance
$d_{\mathcal H}$ (see (\ref{eq:dH})) is dominated by the tangential
distance $d_{\mathcal T}$.
Our main result is as follows.
\begin{theorem}\label{th:main1}
Suppose that\/ $0<c_1\le n_2/n_1\le c_2<\infty$\textup{,} and assume
that the coefficients $b_\ell$ in \textup{(\ref{eq:b-Gamma})}
satisfy some mild technical conditions expressed in terms of the
power series
expansion of the function\/ $y(s)=\ln\left(\sum_\ell b_\ell\mypp
s^\ell\right)$ \textup{(}see more details in Section
\textup{\ref{sec2.1}}\textup{)}. Then for any\/ $\varepsilon>0$
\begin{equation*}
\lim_{n\to\infty}\PP_n\{\varGamma\in\CP_n:\,d_{\mathcal
T}(\tilde\varGamma_n,\gamma^*)\le\varepsilon\}=1.
\end{equation*}
\end{theorem}
\begin{remark}
Universality of the limit shape $\gamma^*$ has its boundaries: as
was shown by Bogachev and Zarbaliev \cite{BZ2,BZ_inverse}, any
$C^3$-smooth, strictly convex curve $\gamma$ started at the origin
may appear as the limit shape with respect to a suitable probability
measure $\PP_n^\gamma$ on $\CP_n$, as $n\to\infty$.
\end{remark}
Like in \cite{BZ4}, our proof employs the elegant probabilistic
approach based on randomization and conditioning (see
\cite{ABT,AT})
first used in the polygonal context by Sinai
\cite{Sinai}. The idea is to introduce a suitable product measure
$\QQ_z$ on the space $\CP=\cup_n\CP_n$ (depending on an auxiliary
``free'' parameter $z=(z_1,z_2)$), such that the measure $\PP_n$ on
$\CP_n$ is recovered as the conditional distribution
$\PP_n(\cdot)=\QQ_z(\cdot\,|\,\CP_n)$. Clearly, this device calls
for the asymptotics of the probability $\QQ_z(\CP_n)$, which is
supplied by proving a suitable local limit theorem. Let us also
point out that the parameter $z$ is calibrated from the asymptotic
equation $\EE_z(\xi_\varGamma)=n\left(1+o(1)\right)$, where
$\xi_\varGamma$ is the right endpoint of the polygonal line
$\varGamma\in\CP$ (so that, e.g.,
$\CP_n=\{\varGamma\in\CP\mynn:\myp\xi_\varGamma=n\}$). The main
novelty that has allowed us to extend and enhance the argumentation
of \cite{BZ4} in a much more general setting considered here is that
we choose to work with cumulants rather than moments (see Section
\ref{sec2.1}), which proves extremely efficient throughout.
\subsubsection*{Layout.}
The rest of the paper is organized as follows. In
Section~\ref{sec2}, we define the families of measures $\QQ_z$ and
$\PP_n$. In Section~\ref{sec3}, suitable values of the parameter
$z=(z_1,z_2)$ are chosen (Theorem~\ref{th:delta12}), which implies
convergence of ``expected'' polygonal lines to the limit curve
$\gamma^*$ (Theorems \ref{th:3.2} and~\ref{th:8.1.1a}). Refined
first-order
moment asymptotics are obtained in Section~\ref{sec4}
(Theorem~\ref{th:4.1}), while higher-order moment sums are analyzed
in Section~\ref{sec5}. Section~\ref{sec7} is devoted to the proof of
the local central limit theorem (Theorem~\ref{th:LCLT}). Finally,
the limit shape result, with respect to both $\QQ_z$ and $\PP_n$, is
proved in Section~\ref{sec8} (Theorems \ref{th:8.2}
and~\ref{th:8.2a}).
\subsubsection*{Some general notations.}
For a row-vector $x=(x_1,x_2)\in \RR^2$, its Euclidean norm (length)
is denoted by $|x|:=(x_1^2+x_2^2)^{1/2}$, and $\langle
x,y\rangle:=x\mypp y^{\myn\topp\!}=x_1y_1+x_2\myp y_2$ is the
corresponding inner product of vectors $x,y\in \RR^2$. We denote
$\ZZ_+:=\{k\in\ZZ: k\ge0\}$, \,$\ZZ_+^2:=\ZZ_+\mynn\times\ZZ_+$\myp,
and similarly $\RR_+:=\{x\in\RR: x\ge0\}$,
$\RR_+^2:=\RR_+\mynn\times\RR_+$\myp.
\section{Probability measures on spaces of convex polygonal lines}
\label{sec2}
\subsection{Global measure\/ \myp$\QQ_{z}$ and conditional
measure\/ $\PP_n$}\label{sec2.1}
Consider the set
\begin{equation}\label{eq:X}
\calX:=\{x=(x_1,x_2)\in\ZZ^2_+:\,\gcd(x_1,x_2)=1\},
\end{equation}
where
``$\gcd$''
stands for
``greatest common divisor''.
Let $\varPhi:= (\ZZ_+)^{\calX}$ be the space
of functions on $\calX$ with nonnegative integer values, and
consider the subspace of functions with \textit{finite support}\/,
$\varPhi_0 :=\{\nu\in\varPhi :\,\#(\supp\myp\nu)<\infty\}$, where $
\supp\myp\nu:=\{x\in \calX:\,\nu(x)>0\}$. It is easy to see that the
space $\varPhi_0$ is in one-to-one correspondence with the space
$\CP=\bigcup_{n\in\ZZ_+^2}\CP_n$ of all (finite) convex lattice
polygonal lines, whereby each $x\in \calX$ determines the
\emph{direction} of a potential edge, only utilized if
$x\in\supp\myp\nu$, in which case the value $\nu(x)>0$ specifies the
\emph{scaling factor}, altogether yielding a vector edge
$x\myp\nu(x)$; finally, assembling all such edges into a polygonal
line is uniquely determined by the fixation of the starting point
(at the origin) and the convexity property.
Let $b_0,b_1,b_2,\dots$ be a sequence of non-negative numbers such
that $b_0>0$ (without loss of generality, we put $b_0=1$) and not
all $b_\ell$'s vanish for $\ell\ge1$, and assume that the
function
\begin{equation}\label{eq:B}
\beta(s):=1+\sum_{\ell=1}^\infty b_\ell\mypp s^\ell
\end{equation}
is finite for $|s|<1$. Let $z=(z_1,z_2)\in (0,1)\times(0,1)$.
Throughout the paper, we shall use the multi-index notation
\begin{equation*}
z^{x}:=z_1^{x_1}z_2^{x_2},\qquad x=(x_1,x_2)\in\ZZ_+^2\myp.
\end{equation*}
Let us now define a probability measure $\QQ_{z}$ on the space
$\varPhi=\ZZ_+^{\calX}$ as the distribution of a random field
$\nu=\{\nu(x)\}_{x\in \calX}$ with mutually independent values and
marginal distributions
\begin{equation}\label{Q}
\QQ_{z}\{\nu(x)=\ell\}=\frac{b_\ell\mypp z^{\ell
x}}{\beta(z^x)}\myp,\qquad \ell=0,1,2,\dots\ \quad (x\in\calX).
\end{equation}
\begin{lemma}\label{pr:F0} For each $z\in (0,1)^2$, the
condition
\begin{equation}\label{N}
\tilde\beta(z):=\prod_{x\in \calX}\beta(z^{x})<\infty
\end{equation}
is necessary and sufficient in order that $\QQ_{z}(\varPhi_0)=1$.
Furthermore, if $\beta(s)$ is finite for all $|s|<1$ then condition
\textup{(\ref{N})} is satisfied.
\end{lemma}
\begin{proof} According to (\ref{Q}),
$\QQ_{z}\{\nu(x)>0\}=1-\beta(z^x)^{-1}$ ($x\in\calX$). Hence,
Borel--Cantelli's lemma implies that $\QQ_{z}\{\nu\in\varPhi_0\}=1$
if and only if $\sum_{x\in\calX}
\bigl(1-\beta(z^x)^{-1}\bigr)<\infty$. In turn, the latter
inequality is equivalent to (\ref{N}).
To prove the second statement,
observe using (\ref{eq:B}) that
\begin{equation}\label{eq:ln-beta0}
\ln\tilde{\beta}(z)=\sum_{x\in\calX}\ln \beta(z^x) \le
\sum_{x\in\calX} \bigl(\beta(z^x)-1\bigr)=\sum_{\ell=1}^\infty
b_\ell \sum_{x\in\calX} z^{\ell x}.
\end{equation}
Furthermore, for any $\ell\ge1$
\begin{align*}
\sum_{x\in \calX}z^{\ell x}&\le \sum_{x_1=1}^\infty z_1^{\ell x_1}+
\sum_{x_1=0}^\infty z_1^{\ell
x_1}\sum_{x_2=1}^\infty z_2^{\ell x_2}\\
&=\frac{z_1^\ell}{1-z_1^\ell}+
\frac{z_2^\ell}{(1-z_1^\ell)(1-z_2^\ell)}\le \frac{z_1^\ell}{1-z_1}+
\frac{z_2^\ell}{(1-z_1)(1-z_2)}\myp.
\end{align*}
Substituting this into (\ref{eq:ln-beta0}) and recalling
(\ref{eq:B}), we obtain
$$
\ln\tilde{\beta}(z^x) \le \frac{\beta(z_1)}{1-z_1}+
\frac{\beta(z_2)}{(1-z_1)(1-z_2)}<\infty,
$$
which implies (\ref{N}).
\end{proof}
Lemma \ref{pr:F0} ensures
that a sample configuration of the
random field $\nu(\cdot)$ belongs ($\QQ_{z}$-a.s.) to the space
$\varPhi_0$ and therefore determines a (random) finite polygonal
line $\varGamma\in \CP$. By the mutual independence of the values
$\nu(x)$, the corresponding $\QQ_z$-probability is given by
\begin{equation}\label{Q1}
\QQ_{z}(\varGamma) =\prod_{x\in \calX}\frac{b_{\nu(x)}\myp
z^{x\myp\nu(x)}}{\beta(z^x)} =\frac{b(\varGamma)\myp z^{\xi}}{\tilde
\beta(z)}\myp,\qquad \varGamma\in\CP,
\end{equation}
where $\xi=\sum_{x\in\calX} x\myp\nu(x)$ is the right endpoint of
$\varGamma$,
and
\begin{equation}\label{eq:b}
b(\varGamma):=\prod_{x\in \calX} \myn b_{\nu(x)}<\infty,\qquad
\varGamma\in\CP.
\end{equation}
\begin{remark}
The infinite product in (\ref{eq:b}) contains only finitely many
terms different from $1$ (since $b_{\nu(x)}=b_0=1$ for
$x\notin\supp\myp\nu$); hence, (\ref{eq:b}) can be rewritten in an
intrinsic form (\ref{eq:b-Gamma}).
\end{remark}
In particular, for the trivial polygonal line
$\varGamma_0\leftrightarrow \nu\equiv0$ formula (\ref{Q1}) yields
\begin{equation*}
\QQ_z(\varGamma_0)=\tilde \beta(z)^{-1}>0.
\end{equation*}
Note, however, that $\QQ_z(\varGamma_0)<1$, since
$\beta(s)>\beta(0)=1$ for $s>0$ and hence, according to definition
(\ref{N}), $\tilde\beta(z)>1$.
On the subspace $\CP_{n}\subset\CP$ of polygonal lines with the
right endpoint fixed at $n=(n_1,n_2)$, the measure $\QQ_{z}$ induces
the conditional distribution
\begin{equation}\label{Pn}
\PP_n(\varGamma):=\QQ_{z}(\varGamma\myp|\myp\CP_n)
=\frac{\QQ_{z}(\varGamma)}{\QQ_{z}(\CP_n)}\myp,\qquad \varGamma\in
\CP_n.
\end{equation}
Formula (\ref{Pn}) is well defined as long as $\QQ_{z}(\CP_n)>0$,
that is, there is at least one polygonal line $\varGamma\in\CP_n$
with $b(\varGamma)>0$ (see (\ref{Q1}) and (\ref{eq:b})). A simple
sufficient condition is as follows.
\begin{lemma}
Suppose that $b_1>0$. Then $\QQ_z(\CP_n)>0$ for all $n\in\ZZ_+^2$
such that $n_1,n_2>0$.
\end{lemma}
\proof Observe that $n=(n_1,n_2)\in\ZZ_+^2$ (with $n_1,n_2\ge1$) can
be represented as
\begin{equation}\label{eq:n=n1+n2}
(n_1,n_2)=(n_1-1,1)+(1,n_2-1),
\end{equation}
where both points $x^{(1)}=(n_1-1,1)$ and $x^{(2)}=(1,n_2-1)$ belong
to the set $\calX$. Moreover, $x^{(1)}\ne x^{(2)}$ unless
$n_1=n_2=2$, in which case instead of (\ref{eq:n=n1+n2}) we can
write $(2,2)=(1,0)+(1,2)$, where again $x^{(1)}=(1,0)\in\calX$,
$x^{(2)}=(1,2)\in\calX$. If $\varGamma^*\in\CP_n$ is a polygonal
line with two edges determined by the values $\nu(x^{(1)})=1$,
$\nu(x^{(2)})=1$ (and $\nu(x)=0$ otherwise), then, according to
definition (\ref{Q1}), $\QQ_z(\CP_n)\ge \QQ_z(\varGamma^*)=b_1^2
z^n\tilde{\beta}(z)^{-1}>0$.
\endproof
The parameter $z$ may be dropped in notation (\ref{Pn}) due to the
following key fact.
\begin{lemma}
The measure\/ $\PP_n$ in\/ \textup{(\ref{Pn})} does not depend on\/
$z$.
\end{lemma}
\begin{proof} If $\CP_n\ni\varGamma \leftrightarrow\nu\in\varPhi_0$ then
$\xi=n$ and hence formula (\ref{Q1}) is reduced to
\begin{equation*}
\QQ_{z}(\varGamma)= \frac{b(\varGamma)\myp z^n}{\tilde
\beta(z)}\myp,\qquad \varGamma\in \CP_n.
\end{equation*}
Accordingly, using (\ref{N}) and (\ref{Pn}) we get the expression
\begin{equation}\label{condP}
\PP_n(\varGamma)=\frac{b(\varGamma)}{\sum_{\varGamma'\myn\in\CP_n}\mynn
b(\varGamma')}\myp,\qquad \varGamma\in \CP_n,
\end{equation}
which is $z$-free.
\end{proof}
\subsection{A class of measures\/ $\QQ_{z}$}\label{sec2.2}
Recalling expansion (\ref{eq:B}) for the generating function
$\beta(s)$ (with $\beta(0)=b_0=1$), consider the corresponding
expansion of its logarithm,
\begin{equation}\label{eq:ln-beta}
\ln\beta(s)=\sum_{k=1}^\infty a_k\myp s^k,\qquad |s|<1.
\end{equation}
\begin{remark}\label{rm:b=a}
Substituting expansion (\ref{eq:B}) into (\ref{eq:ln-beta}) it is
clear that $a_1=b_1$; more generally, if $\ell^*:=\min\{\ell\ge 1:
b_\ell>0\}$ and $k^*:=\min\{k\ge 1: a_k\ne0\}$ then $\ell^*=k^*$ and
$b_{\ell^*}=a_{k^*}$.
\end{remark}
Under the measure $\QQ_z$ defined in (\ref{Q}), the probability
generating function $\phi_{\nu}(s;x):=E_z [s^{\nu(x)}]$ of $\nu(x)$
is given by the ratio
\begin{equation}\label{eq:pgf}
\phi_{\nu}(s;x)=\frac{\beta(sz^x)}{\beta(z^x)},\qquad |s|\le1
\end{equation}
(for notational simplicity, we suppress the dependence on $z$, which
should cause no confusion), and so its logarithm is expanded as
\begin{align}\label{eq:pgf-T}
\ln \phi_{\nu}(s;x)&= \ln\beta(sz^x)-\ln\beta(z^x) =
\sum_{k=1}^\infty a_k(s^k-1)\mypp z^{kx}, \qquad |s|\le1.
\end{align}
Likewise, the characteristic function
$\varphi_{\nu}(t;x):=\EE_z[\rme^{\myp\rmi t\nu(x)}]$ is given by
\begin{equation}\label{eq:c.f.}
\varphi_{\nu}(t;x)=\frac{\beta(z^x\rme^{\myp\rmi t})}{\beta(z^x)},
\qquad t\in\RR.
\end{equation}
and the principal branch of its logarithm (corresponding to $\ln
\varphi_{\nu}(0;x)=0$) is represented as
\begin{equation}\label{eq:ln_c.f.}
\ln \varphi_{\nu}(t;x)=\sum_{k=1}^\infty a_k(\rme^{\myp\rmi
kt}-1)\mypp z^{kx},\qquad t\in\RR.
\end{equation}
For $q\in\NN$, denote by $m_q=m_q(x):=\EE_{z}[\nu(x)^q]$ the moments
of $\nu(x)$, and let $\varkappa_q=\varkappa_q(x)$ be the cumulants
of $\nu(x)$, with the exponential generating function
\begin{equation}\label{eq:varkappa}
\ln \phi_{\nu}(\rme^t;x)=\sum_{q=1}^\infty
\varkappa_q(x)\mypp\frac{t^q}{q!}\myp.
\end{equation}
Substituting (\ref{eq:pgf-T}) into (\ref{eq:varkappa}) and Taylor
expanding the exponential function, we get
$$
\ln \phi_{\nu}(\rme^t;x)=\sum_{k=1}^\infty a_k(\rme^{kt}-1)\mypp
z^{kx}=\sum_{q=1}^\infty \frac{t^q}{q!}\sum_{k=1}^\infty k^q a_k
z^{kx},
$$
and by a comparison with (\ref{eq:varkappa}) it follows that
\begin{equation}\label{eq:cumulants}
\varkappa_q(x)=\sum_{k=1}^\infty k^q a_k z^{kx},\qquad q\in\NN.
\end{equation}
In particular, from (\ref{eq:cumulants}) we obtain the mean and
variance of $\nu(x)$,
\begin{gather}\label{eq:nu}
\EE_{z}[\nu(x)]=m_1(x)=\varkappa_1(x)=\sum_{k=1}^\infty
k a_{k}\myp z^{kx},\\[-.3pc]
\label{eq:nu2}
\Var[\nu(x)]=m_2(x)-m_1(x)^2=\varkappa_2(x)=\sum_{k=1}^\infty k^2
a_{k}\mypp z^{kx}.
\end{gather}
More generally, using a well-known recursion between the cumulants
and moments (see, e.g., \cite[\S\myp3.14]{KS})
\begin{equation*}
m_q=\varkappa_q+\sum_{i=1}^{q-1}\binom{q-1}{i-1}\,\varkappa_i
\,m_{q-i}
\end{equation*}
it is easy to see by a simple induction that the moments $m_q$
($q\in\NN$) are expressed as linear combinations of the cumulants
$\varkappa_1,\dots,\varkappa_q$ with positive (in fact, integer)
coefficients, which gives, in view of (\ref{eq:cumulants}),
\begin{equation}\label{eq:m=varkappa+}
m_q(x)=\varkappa_q(x)+\sum_{i=1}^{q-1} C_{i,\myp
q}\myp\varkappa_i(x)=\varkappa_q(x)+\sum_{i=1}^{q-1} C_{i,\myp
q}\sum_{k=1}^\infty k^{i}
a_k z^{kx},
\end{equation}
with $C_{i,\myp q}>0$ \,($i=1,\dots,q-1$).
Furthermore, using a rescaling relation
$\varkappa_q[cX]=c^q\varkappa_q[X]$ and the additive property of
cumulants for independent summands, we obtain the cumulants of the
random variables $\xi_j=\sum_{x\in\mathcal{X}}x_j\myp\nu(x)$
\,($j=1,2$),
\begin{equation}\label{eq:cumulants-xi}
\varkappa_q[\xi_j]=\sum_{x\in\mathcal{X}}x_j^q
\myp\varkappa_q(x)=\sum_{x\in\mathcal{X}}x_j^q \sum_{k=1}^\infty k^q
a_k z^{kx},
\end{equation}
and, similarly to (\ref{eq:m=varkappa+}), the corresponding moments
\begin{equation}\label{eq:moments-xi}
\EE_z(\xi_j^q)=\varkappa_q[\xi_j]+\sum_{i=1}^{q-1} C_{i,\myp
q}\sum_{x\in\mathcal{X}}x_j^{\myp i}
\sum_{k=1}^\infty k^{i}
a_k z^{kx}.
\end{equation}
For $s\in\CC$ such that $\sigma:=\Re s>0$, denote
\begin{equation}\label{eq:A2-}
A(s):=\sum_{k=1}^\infty \frac{a_{k}}{k^s}\myp,\qquad
A^+(\sigma):=\sum_{k=1}^\infty \frac{|a_{k}|}{k^\sigma}\le\infty.
\end{equation}
Most of our results are valid under the condition $A^+(2)<\infty$,
or sometimes $A^+(1)<\infty$ (in particular, in Theorem
\ref{th:4.1}). However, for a local limit theorem (see Theorem
\ref{th:LCLT}) we require an additional technical condition on the
generating function $\beta(s)$.
\begin{assumption}\label{as:7.1}
The coefficients $(a_k)$ in
expansion \textup{(\ref{eq:B})} of\/ $\ln\beta(s)$ are such that
$a_1>0$ and, for any $\theta\in(0,1)$ and all $t\in\RR$, the
following inequality holds, with some constant $C_1>0$,
\begin{equation}\label{eq:<a1}
\sum_{k=1}^\infty a_k \theta^k(1-\cos kt)\ge C_1 a_1\theta(1-\cos
t).
\end{equation}
\end{assumption}
\begin{remark}\label{rm:a1>0}
Assumption \ref{as:7.1} is obviously satisfied (with $C_1=1$) when
\textit{all $a_k$ are positive}.
\end{remark}
Due to Remark \ref{rm:b=a}, the condition $a_1>0$ is equivalent to
$b_1>0$. Moreover, from (\ref{eq:c.f.}) and (\ref{eq:ln_c.f.}) we
note that
\begin{align}
\ln|\varphi_{\nu}(t;x)|=\frac12\ln\frac{\beta(z^x\rme^{\myp\rmi
t})\beta(z^x\rme^{-\rmi t})}{\beta(z^x)^2} \label{eq:|phi|}
=-\sum_{k=1}^\infty a_k z^{kx}\bigl(1-\cos kt\bigr);
\end{align}
hence, condition (\ref{eq:<a1}) can be equivalently rewritten (for
any $\theta\in(0,1)$ and all $t\in \RR$) as
\begin{equation}\label{eq:<a1-1}
\frac12\ln\frac{\beta(\theta\mypp\rme^{\myp\rmi
t})\myp\beta(\theta\mypp\rme^{-\rmi t})}{\beta(\theta)^2} \le -C_1
b_1 \theta\mypp(1-\cos t).
\end{equation}
\subsection{Examples}\label{sec2.3}
Let us now consider a few illustrative examples. The first three
have direct analogues in the theory of (one-dimensional)
decomposable combinatorial structures, corresponding, respectively,
to the three well-known meta-classes: multisets, selections and
assemblies (see
\cite{ABT,AT,GSE}).
To the best of our knowledge, Example \ref{ex:4}
was first considered in \cite{Bogachev} in the
context of integer partitions.
\begin{example}[multisets]\label{ex:1}
For $r\in(0,\infty)$, $\rho\in(0,1]$, let $\QQ_{z}$ be a measure
determined by formula (\ref{Q}) with coefficients (\ref{eq:b-k-r}).
A particular case with $\rho=1$ was considered in \cite{BZ4}. Note
that $b_0=1$, in accordance with our convention in
Section~\ref{sec2.1}, and $b_1=r\rho>0$. By the binomial expansion
formula, the generating function of sequence (\ref{eq:b-k-r}) is
given by
\begin{equation}\label{G1}
\beta(s)=(1-\rho s)^{-r}, \qquad |s|<\rho^{-1},
\end{equation}
and formula (\ref{Q}) specializes to
\begin{equation}\label{Qb1}
\QQ_{z}\{\nu(x)=\ell\}=\binom{r+\ell-1}{\ell}\, \rho^\ell z^{\ell
x}(1-\rho z^x)^r,\ \ \quad \ell\in\ZZ_+,
\end{equation}
which is a negative binomial distribution with parameters $r$ and
$p=1-\rho z^x$.
If $r=1$ then $b_\ell=\rho^\ell$, \,$\beta(s)=(1-\rho s)^{-1}$ and,
according to (\ref{Qb1}),
\begin{equation*}
\QQ_{z}\{\nu(x)=\ell\}= \rho^\ell z^{\ell x}(1-\rho z^x),\ \ \quad
\ell\in\ZZ_+.
\end{equation*}
In turn, from formulas (\ref{eq:b-Gamma}) and (\ref{condP}) we get
\begin{equation}\label{eq:r=1}
\PP_n(\varGamma)=\frac{\rho^{N_{\varGamma}}}{\sum_{\varGamma'\in\CP_n}
\rho^{N_{\varGamma'}}},\qquad \varGamma\in\CP_n,
\end{equation}
where $N_{\varGamma}:= \sum_{x\in \calX} \nu(x)$ is the total number
of integer points on $\varGamma\setminus\{0\}$.
Furthermore, if also $\rho=1$ then (\ref{eq:r=1}) is reduced to the
uniform distribution on $\CP_n$ (see (\ref{condP})),
$$
\PP_n(\varGamma)=\frac{1}{\#(\CP_n)}\myp,\qquad \varGamma\in\CP_n.
$$
In the general case, using (\ref{G1}) we note that
$$
\ln \beta(s)=-r\ln(1-\rho s)=r\sum_{k=1}^\infty\frac{\rho^k
s^k}{k}\myp,
$$
and so the coefficients $(a_k)$ in expansion (\ref{eq:ln-beta}) are
given by
$$
a_k=\frac{r\rho^k}{k}>0,\qquad k\in\NN.
$$
As pointed out in Remark \ref{rm:a1>0}, this implies that Assumption
\ref{as:7.1} is satisfied; also, it readily follows that
$A^+(\sigma)<\infty$ for any $\sigma>0$.
\end{example}
\begin{example}[selections]\label{ex:2}
For \,$r\in\NN$, \,$\rho\in(0,1]$, consider the generating function
\begin{equation}\label{eq:f(s)2}
\beta(s)=(1+\rho s)^r,\qquad |s|<\rho^{-1},
\end{equation}
with the coefficients in expansion (\ref{eq:B}) given by
\begin{equation}\label{b_k2}
b_\ell=\binom{r}{\ell}\,\rho^\ell=\frac{r(r-1)\cdots(r-\ell+1)}{\ell!}\,\rho^\ell,\qquad
\ell=0,1,\dots,r.
\end{equation}
In particular, $b_0=1$, $b_1=r\rho>0$. Accordingly, formula
(\ref{Q}) gives a binomial distribution
\begin{equation}\label{Qb2}
\QQ_{z}\{\nu(x)=\ell\}=\binom{r}{\ell} \frac{\rho^\ell z^{\ell
x}}{(1+\rho z^x)^{r}},\ \ \quad \ell=0,1,\dots,r,
\end{equation}
with parameters $r$ and $p=\rho z^x(1+\rho z^x)^{-1}$. From
(\ref{eq:f(s)2}) we obtain
$$
\ln \beta(s)=r\ln (1+\rho s)=r\sum_{k=1}^\infty
\frac{(-1)^{k-1}\rho^k}{k}\,s^k,
$$
hence the coefficients $(a_k)$ in expansion (\ref{eq:ln-beta}) are
given by
$$
a_k=\frac{r(-1)^{k-1}\rho^k}{k},\qquad k\in\NN,
$$
and in particular $a_1=r\rho>0$. Note that $A^+(\sigma)<\infty$ for
any $\sigma>0$.
In the special case $r=1$, the measure $\QQ_z$ is concentrated on
the subspace $\check{\CP}$ of polygonal lines with ``simple'' edges,
that is, containing no lattice points between the adjacent vertices.
Here we have $b_0=1$, $b_1=\rho$ and $b_\ell=0$ ($\ell\ge2$), so
that (\ref{Qb2}) is reduced to
\begin{equation*}
\QQ_{z}\{\nu(x)=\ell\}= \frac{\rho^\ell z^{\ell x}}{1+\rho z^x},\ \
\quad \ell=0,1\ \ \quad (x\in\calX),
\end{equation*}
Accordingly, formula (\ref{condP}) specifies on the corresponding
subspace $\check{\CP}_n$ the distribution
\begin{equation}\label{eq:r=1-2}
\PP_n(\varGamma)=\frac{\rho^{N_{\varGamma}}}{\sum_{\varGamma'\in\check\CP_n}
\rho^{N_{\varGamma'}}}\myp,\qquad \varGamma\in \check{\CP}_n,
\end{equation}
where the number of integer points $N_\varGamma$ coincides here with
the number of vertices on $\varGamma\setminus\{0\}$. Furthermore, if
also $\rho=1$ then (\ref{eq:r=1-2}) is reduced to the uniform
distribution on $\check{\CP}_n$,
\begin{equation*}
\PP_n(\varGamma)=\frac{1}{\# (\check{\varPi}_{n})}\myp,\qquad
\varGamma\in \check{\CP}_n.
\end{equation*}
Finally, let us check that Assumption \ref{as:7.1} holds (with
$C_1=(1+\rho)^{-2}$). It
is more convenient to use
version (\ref{eq:<a1-1}). Substituting
(\ref{eq:f(s)2})
and recalling that $b_1=r\rho>0$, we obtain
\begin{align*}
\frac12\ln\left(\frac{\beta(\theta\mypp\rme^{\myp\rmi
t})\myp\beta(\theta\mypp\rme^{-\rmi t})}{\beta(\theta)^2}\right)
&=\frac{r}{2}\ln\left(\frac{1+2\rho\myp\theta\cos t+\rho^2\theta^2}{(1+\rho\myp\theta)^2}\right)\\
&\le \frac{r}{2}\left(\frac{1+2\rho\myp\theta\cos t+
\rho^2\theta^2}{(1+\rho\myp\theta)^2}-1\right)
\le -\frac{b_1 \theta\mypp(1-\cos t)}{(1+\rho)^2}.
\end{align*}
\end{example}
\begin{example}[assemblies]\label{ex:3}
For $r\in(0,\infty)$, \,$\rho\in[0,1]$, consider the generating
function
\begin{equation}\label{eq:beta3}
\beta(s)=\exp\left(\frac{r s}{1-\rho
s}\right)=\exp\left(r\sum_{k=1}^\infty s^{k}\rho^{k-1}\right),\qquad
|s|<\rho^{-1}.
\end{equation}
Clearly, the corresponding coefficients $b_\ell$ in expansion
(\ref{eq:B}) are positive, with $b_0=1$, $b_1=r$, $b_2=\frac12\myp
r^2+r\myn\rho$, etc.; more systematically, one can use the
well-known Fa\`a di Bruno's formula generalizing the chain rule to
higher derivatives (see, e.g., \cite[Ch.\,I, \S12, p.\,34]{Jordan})
to obtain
\begin{equation}\label{eq:Faa}
b_\ell=\rho^\ell\sum_{m=1}^\ell\left(\frac{r}{\rho}\right)^m\!\!\sum_{(j_1\myn,\dots,\myp
j_\ell) \myp\in\myp\mathcal{J}_m}\frac{1}{j_1\myn!\cdots j_\ell
!}\myp, \qquad \ell\in\NN,
\end{equation}
where $\mathcal{J}_m$ is the set of all non-negative integer
$\ell$-tuples $(j_1,\dots,j_\ell)$ such that $j_1+\dots+j_\ell=m$
and $1\cdot j_1+2\cdot j_2+\dots+\ell\cdot j_\ell =\ell$.
\begin{remark}
Note that the $\ell$-tuples $(j_1,\dots,j_\ell)\in\mathcal{J}_m$ are
in a one-to-one correspondence with partitions of $\ell$ involving
precisely $m$ different integers as parts, where an element $j_i$
has the meaning of the multiplicity of $i\in\NN$ (i.e., the number
of times $i$ is used in a partition of $\ell$).
\end{remark}
Taking the logarithm of (\ref{eq:beta3}), we see that the
coefficients $a_k$ in (\ref{eq:ln-beta}) are given by
\begin{equation}\label{eq:ak}
a_k=r\rho^{k-1}>0,\qquad k\in\NN.
\end{equation}
Therefore, Assumption \ref{as:7.1} is automatic; moreover,
$A^+(\sigma)<\infty$ for any $\sigma>0$, except for the case
$\rho=1$ where $A^+(\sigma)<\infty$ only for $\sigma>1$.
In the particular case $\rho=0$, we have $\beta(s)=\rme^{rs}$ and so
expression (\ref{eq:Faa}) is replaced by $b_\ell=r^\ell/\ell!$,
whereas (\ref{eq:ak}) simplifies to $a_1=r$ and $a_k=0$ for $k\ge2$.
The random variables $\nu(x)$ have a Poisson distribution with
parameter $r\myn z^x$,
\begin{equation*}
\QQ_{z}\{\nu(x)=\ell\}=\frac{r^\ell z^{\ell x}}{\ell!}\,\rme^{-r
z^x},\ \ \quad \ell \in\ZZ_+\myp,
\end{equation*}
which leads, according to (\ref{condP}), to the following
distribution on $\CP_n$
\begin{equation*}
\PP_n(\varGamma)=\left(\sum_{\{\ell^{\myp\prime}_x\}\in\CP_n}
\prod_{x\in\calX}\frac{r^{\ell^{\myp\prime}_x}}{\ell^{\myp\prime}_x!}\right)^{-1}
\prod_{x\in\calX}\frac{r^{\ell_x}}{\ell_x!}\myp,\qquad
\varGamma\leftrightarrow\{\ell_x\}\in \CP_n.
\end{equation*}
\end{example}
\begin{example}\label{ex:4}
Let $r\in(0,\infty)$, \,$\rho\in(0,1]$, and consider the generating
function
\begin{equation}\label{eq:f(s)4}
\beta(s)=\left(\frac{-\ln(1-\rho s)}{\rho
s}\right)^r=\left(1+\sum_{\ell =1}^\infty \frac{\rho^\ell s^{\ell
}}{\ell +1}\right)^r=:f(s)^r.
\end{equation}
From (\ref{eq:f(s)4}) it is clear that $b_0=1$, $b_1=\frac12\mypp
r\rho>0$ and, more generally, all $b_\ell>0$. Let us analyze the
coefficients $(a_k)$ in the power series expansion of
$\ln\beta(s)=r\ln f(s)$ (see (\ref{eq:ln-beta})). Differentiation of
this identity with respect to $s$ gives
\begin{equation}\label{eq:dif-once}
r f'(s)=f(s)\sum_{k=1}^\infty ka_ks^{k-1}.
\end{equation}
Differentiating (\ref{eq:dif-once}) further $m$ times ($m\ge0$), by
the Leibniz rule we obtain
$$
f^{(m+1)}(s)=\frac{1}{r}\sum_{j=0}^m \binom{m}{j}f^{(m-j)}(s)\sum_{k=j+1}^\infty
\frac{k!}{(k-j-1)!} a_ks^{k-j-1},
$$
and in particular
\begin{equation}\label{eq:f'(0)}
f^{(m+1)}(0)=\frac{1}{r}\sum_{j=0}^m \binom{m}{j}f^{(m-j)}(0)
(m+1)! \,a_{m+1}.
\end{equation}
But we know from (\ref{eq:f(s)4}) that $ f^{(j)}(0)=\rho^j
j!/(j+1)$, so (\ref{eq:f'(0)}) specializes to the equation
$$
\frac{\rho^{m+1}(m+1)!}{m+2}=\frac{1}{r}\sum_{j=0}^m
\frac{m!}{j!\myp(m-j)!} \cdot\frac{\rho^{m-j}(m-j)!}{m-j+1}(j+1)!
\,a_{j+1},
$$
or, after some cancellations,
\begin{equation}\label{eq:m/m}
\frac{m+1}{m+2}=\frac{1}{r}\sum_{j=0}^m \frac{\rho^{-j-1}
(j+1)}{m-j+1} \,a_{j+1}.
\end{equation}
Denoting for short $\tilde{a}_j:=r^{-1}\rho^{-j}j a_j$, equation
(\ref{eq:m/m}) simplifies to
\begin{equation}\label{eq:a_m+1}
\frac{m+1}{m+2}=\frac{\tilde{a}_1}{m+1}+\frac{\tilde{a}_2}{m}+\dots+\frac{\tilde{a}_{m}}{2}+
\tilde{a}_{m+1}\myp.
\end{equation}
Setting here $m=0,1,2,3,\dots$ we can in principle find successively
all $\tilde{a}_m$,
$$
\tilde{a}_1=\frac{1}{2},\quad \tilde{a}_2=\frac{5}{12},\quad
\tilde{a}_3=\frac{3}{8},\quad \tilde{a}_4=\frac{251}{720},\ \dots,
$$
but the fractions quickly become quite cumbersome. However, it is
not hard to obtain suitable estimates of $\tilde{a}_m$. Observe that
(\ref{eq:a_m+1}) implies
\begin{align*}
\frac{m+1}{m+2} &\le
\frac{\tilde{a}_1}{m}+\frac{\tilde{a}_2}{m-1}+\dots+\tilde{a}_{m}+
\tilde{a}_{m+1}=\frac{m}{m+1}+\tilde{a}_{m+1},
\end{align*}
and it follows that
$$
\tilde{a}_{m+1}\ge
\frac{m+1}{m+2}-\frac{m}{m+1}=\frac{1}{(m+1)(m+2)}>0,
$$
or explicitly
\begin{equation}\label{eq:a>}
a_{m+1}\ge \frac{r\rho^{m+1}}{(m+1)^2(m+2)}>0.
\end{equation}
On the other hand, from (\ref{eq:a_m+1}) we get
$$
\tilde{a}_{m+1}=\frac{m+1}{m+2}-\frac{\tilde{a}_1}{m+1}-\frac{\tilde{a}_2}{m}-\dots-\frac{\tilde{a}_{m}}{2}\le
\frac{m+1}{m+2}-\frac{\tilde{a}_1}{m+1}\myp,
$$
hence
$$
\tilde{a}_{m+1}\le
\frac{m+1}{m+2}-\frac{1/2}{m+1}=\frac{2m^2+3m}{2(m+1)(m+2)}
$$
and therefore
\begin{equation}\label{eq:a<}
a_{m+1}\le \frac{r\rho^{m+1}(2m^2+3m)}{2(m+1)^2(m+2)}\myp.
\end{equation}
As a result, combining (\ref{eq:a>}) and (\ref{eq:a<}) we obtain,
for all $k\in\NN$,
$$
\frac{r\rho^k}{k^2(k+1)}\le a_k\le
\frac{r\rho^k(2k^2-k-1)}{2k^2(k+1)}\le\frac{r\rho^k}{k+1}\myp.
$$
In particular, this implies that $ A^+(\sigma)<\infty$ for any
$\sigma>0$; furthermore, since all $a_k>0$ it follows that
Assumption \ref{as:7.1} is automatically satisfied.
\end{example}
\begin{remark}
Specific choices of the coefficients $(b_\ell)$ in Examples
\ref{ex:1}--\ref{ex:4} above can be used in the context of integer
partitions (see, e.g.,
\cite{GSE,V3,Yakubovich}
and also a recent preprint \cite{Bogachev}).
More specifically, Example \ref{ex:1} corresponds to the ensemble of
weighted partitions including the case of all unrestricted
partitions under the uniform distribution; Example \ref{ex:2} leads
to (weighted) partitions with bounds on the multiplicities of parts,
including the case of uniform partitions with distinct parts;
Example \ref{ex:3} corresponds to partitions representing the cycle
structure of permutations; finally, Example \ref{ex:4}
introduced in \cite{Bogachev}
defines a new ensemble of random partitions. Note that the limit
shapes of partitions (or rather their Young diagrams) in the first
three cases are known to exist, at least under some technical
conditions on the coefficients (see
\cite{Bogachev,EG,V3,Yakubovich},
but they are all drastically different from each other, as opposed
to the case of lattice polygonal lines representing strict vector
partitions, for which the limit shape is universal in all four
examples.
\end{remark}
\section{Asymptotics of the expectation}
\label{sec3}
In what follows, the asymptotic notation of the form $x_n\asymp y_n$
with $n=(n_1,n_2)$ means that
\begin{equation*}
0<\liminf_{n_1\myn,\myp
n_2\to\infty}\frac{x_n}{y_n}\le\limsup_{n_1\myn,\myp n_2\to\infty}
\frac{x_n}{y_n}<\infty.
\end{equation*}
We also use the standard notation $x_n\sim y_n$ for $x_n/y_n\to1$ as
$n_1,n_2\to\infty$.
Throughout the paper, we adopt the following convention about the
limit $n\to\infty$.
\begin{assumption}\label{as:c}
The notation $n\to\infty$ signifies that $n_1,n_2\to\infty$ in such
a way that $n_1\asymp n_2$. In particular, this implies that
$n_1\asymp |n|$, $n_2\asymp |n|$, where
$|n|=(n_1^2+n_2^2)^{1/2}\to\infty$.
\end{assumption}
\subsection{Calibration of the parameter $z$}\label{sec3.1}
We want to find the parameter $z=(z_1,z_2)$ from the asymptotic
conditions
\begin{equation}\label{calibr1}
\EE_{z}(\xi_1)\sim n_1,\qquad \EE_{z}(\xi_2)\sim n_2\qquad
(n\to\infty),
\end{equation}
where $\xi_j=\sum_{x\in\calX}x_j\myp\nu(x)$ and $\EE_{z}$ denotes
expectation with respect to $\QQ_{z}$. Set
\begin{equation}\label{alpha}
z_j=\rme^{-\alpha_j},\qquad \alpha_j=\delta_j\mypp n_j^{-1/3}\qquad
(j=1,2),
\end{equation}
where the quantities $\delta_1,\delta_2>0$ (possibly depending on
the ratio $n_2/n_1$) are presumed to be bounded from above and
separated from zero. Hence, recalling
formula (\ref{eq:nu}), we get
\begin{equation}\label{E_i'}
\EE_{z}(\xi)=\sum_{k=1}^\infty k a_{k} \sum_{x\in\calX}x\mypp
\rme^{-k\langle\alpha,x\rangle}.
\end{equation}
To deal with sums over the set $\calX$, the following lemma will be
instrumental. Recall that the \textit{M\"obius function} $\mu(m)$
($m\in\NN$) is defined as follows: $\mu(1):=1$, \,$\mu(m):=(-1)^{d}$
if $m$ is a product of $d$ different prime numbers, and $\mu(m):=0$
otherwise (see \cite[\S16.3, p.\:234]{HW}); in particular,
$|\mu(m)|\le 1$ for all $m\in\NN$.
\begin{lemma}\label{lm:Mobius}
Suppose that a function $f:\RR_+^2\to\RR$ is such that $f(0,0)=0$
and\textup{,} for any $h>0$\textup{,}
\begin{equation}\label{eq:sum_sum<}
\sum_{k=1}^\infty\sum_{x\in\ZZ_+^2} |f(hkx)|<\infty.
\end{equation}
For $h>0$, consider the functions
\begin{gather}\label{f+lemma}
F^\sharp(h):=\sum_{x\in \calX} f(hx),\\
\label{eq:F_lemma} F(h):=\sum_{m=1}^\infty F^\sharp(hm)
=\sum_{m=1}^\infty\sum_{x\in \calX} f(hmx).
\end{gather}
Then the following identities hold for all $h>0$
\begin{align}\label{F_lemma}
F(h)&= \sum_{x\in\ZZ^2_+}f(hx),\\
\label{eq:Mobius} F^\sharp(h)&=\sum_{m=1}^\infty\mu(m)F(hm),
\end{align}
where $\mu(m)$ is the M\"obius function.
\end{lemma}
\begin{proof}
Recalling definition (\ref{eq:X}) of the set $\calX$, observe that
$\ZZ^2_+=\bigsqcup_{\myp m=0}^{\myp\infty} m\calX$; hence,
(\ref{eq:F_lemma}) is reduced to (\ref{F_lemma}). Then
representation (\ref{eq:Mobius}) follows from the M\"obius inversion
formula (see \cite[Theorem~270, p.\:237]{HW}), provided that
$\sum_{k,\myp{}m} |F^\sharp(hkm)|<\infty$. To verify the last
condition, using (\ref{f+lemma}) we obtain (cf.\ (\ref{eq:F_lemma})
and (\ref{F_lemma}))
\begin{align*}
\sum_{k,\myp{}m=1}^\infty |F^\sharp(kmh)|&\le \sum_{k=1}^\infty
\left(\sum_{m=1}^\infty\sum_{x\in \calX} |f(hkmx)|\right)
=\sum_{k=1}^\infty \sum_{x\in \ZZ^2_+} |f(hkx)|<\infty,
\end{align*}
according to (\ref{eq:sum_sum<}). This completes the proof.
\end{proof}
\begin{theorem}\label{th:delta12}
Suppose that $A^+(2)<\infty$ \textup{(}see
\textup{(\ref{eq:A2-})}\textup{),} and choose $\delta_1,\delta_2$
in\/ \textup{(\ref{alpha})} as follows\textup{,}
\begin{equation}\label{delta12}
\delta_1=\kappa\myp (n_2/n_1)^{1/3},\qquad \delta_2=\kappa\myp
(n_1/n_2)^{1/3},
\end{equation}
where
\begin{equation}\label{eq:kappa}
\kappa\myn:=\left(\frac{A(2)}{\zeta(2)}\right)^{\myn1/3}
\end{equation}
and $\zeta(2):=\sum_{k=1}^{\vphantom{y}\infty}\myn{}
k^{-2}=\pi^2/6$. Then conditions\/ \textup{(\ref{calibr1})} are
satisfied.
\end{theorem}
\begin{remark}
Observe that (\ref{alpha}) and (\ref{delta12}) imply the scaling
relations
\begin{equation}\label{eq:alpha2:1}
\alpha^2_1\alpha_2\mypp n_1= \alpha_1\alpha^2_2\mypp
n_2=\kappa^3,\qquad \alpha_2\myp n_2=\alpha_1 n_1.
\end{equation}
\end{remark}
\begin{proof}[Proof of Theorem \textup{\ref{th:delta12}}]
Let us prove (\ref{calibr1}) for $\xi_1$ (the proof for $\xi_2$ is
similar). Setting
\begin{equation}\label{f+}
f(x):=x_1 \myp\rme^{-\langle \alpha,x\rangle},\qquad
x\in\RR^2_+\myp,
\end{equation}
and following notations (\ref{f+lemma}) and (\ref{eq:F_lemma}) of
Lemma \ref{lm:Mobius}, a projection of equation (\ref{E_i'}) to the
first coordinate takes the form
\begin{equation}\label{E_1F}
\EE_{z}(\xi_1) =\sum_{k=1}^\infty a_{k} F^\sharp(k).
\end{equation}
Note that
\begin{equation}\label{F}
F(h) =h\sum_{x_1=1}^\infty
x_1\rme^{-h\alpha_1x_1}\sum_{x_2=0}^\infty \rme^{-h\alpha_2x_2}
=\frac{h\mypp
\rme^{-h\alpha_1}}{(1-\rme^{-h\alpha_1})^2(1-\rme^{-h\alpha_2})}\myp,
\end{equation}
and it easily follows that condition (\ref{eq:sum_sum<}) of Lemma
\ref{lm:Mobius} is satisfied. Hence, using (\ref{eq:Mobius}) and
(\ref{F}), we can rewrite (\ref{E_1F}) as
\begin{equation}\label{E_1}
\EE_{z}(\xi_1)= \sum_{k=1}^\infty a_{k}\sum_{m=1}^\infty \mu(m) \myp
F(km) = \sum_{k=1}^\infty k a_{k} \sum_{m=1}^\infty
\frac{m\myp\mu(m)\mypp \rme^{-km\alpha_1}}
{(1-\rme^{-km\alpha_1\myn})^2\myp(1-\rme^{-km\alpha_2\myn})}\myp,
\end{equation}
or, recalling relations (\ref{eq:alpha2:1}),
\begin{equation}\label{E_1'}
n_1^{-1}\EE_{z}(\xi_1)=\frac{1}{\kappa^3} \sum_{k,\myp m=1}^\infty k
a_{k}\mypp m\myp\mu(m) \,\frac{\alpha_1^2\alpha_2 \mypp
\rme^{-km\alpha_1}}
{(1-\rme^{-km\alpha_1\myn})^2\myp(1-\rme^{-km\alpha_2\myn})}\myp.
\end{equation}
Note that for any $b>0$, $\theta>0$, there is a global bound
\begin{equation}\label{eq:exp_bound}
\frac{\rme^{-\theta\myp t}}{(1-\rme^{-t})^b} \le C\mypp
t^{-b},\qquad t>0,
\end{equation}
with some constant $C=C(b,\theta)>0$. This gives, uniformly in $k$
and $m$,
\begin{equation}\label{eq:2,1}
\frac{\rme^{-km\alpha_1/2}}
{(1-\rme^{-km\alpha_1\myn})^2}=\frac{O(1)}{(km\alpha_1)^{2}}\myp,\qquad
\frac{\rme^{-(n_2/n_1)\mypp km\alpha_2/2}}
{1-\rme^{-km\alpha_2\myn}}=\frac{O(1)}{km\alpha_2}\myp,
\end{equation}
where in the second estimate we used Assumption \ref{as:c}.
Therefore, the summand in (\ref{E_1'}) is bounded by
$O\bigl(|a_k|\mypp k^{-2}m^{-2}\bigr)$, which is a term of a
convergent series due to the assumption $A^+(2)<\infty$. Hence, by
Lebesgue's dominated convergence theorem
we obtain
\begin{equation}\label{zeta32}
\lim_{n\to\infty} n_1^{-1} \EE_{z}(\xi_1)=
\frac{1}{\kappa^3}\sum_{k=1}^\infty
\frac{a_{k}}{k^2}\sum_{m=1}^\infty
\frac{\mu(m)}{m^2}=\frac{A(2)}{\kappa^3\zeta(2)}=1,
\end{equation}
according to (\ref{eq:kappa}); we also used the identity
\begin{equation}\label{eq:zeta^{-1}}
\sum_{m=1}^\infty \frac{\mu(m)}{m^s}=\frac{1}{\zeta(s)}\myp,
\end{equation}
which readily follows by the M\"obius inversion formula
(\ref{eq:Mobius}) applied to $F^\sharp(h)=h^{-s}$,
\,$F(h)=\sum_{m=1}^\infty (hm)^{-s} \allowbreak =h^{-s}\myp\zeta(s)$
\,(cf.\ \cite[Theorem~287, p.\;250]{HW}).
\end{proof}
\begin{assumption}\label{as:z}
Throughout the rest of the paper, we assume that the parameters
$z_1,z_2$ are chosen according to formulas (\ref{alpha}),
(\ref{delta12}). In particular, the measure $\QQ_z$ becomes
dependent on $n=(n_1,n_2)$, as well as the $\QQ_z$-probabilities and
the corresponding expected values.
\end{assumption}
\subsection{Asymptotics of the mean polygonal lines}\label{sec3.2}
For $\varGamma\in\CP$, denote by $\varGamma(t)$ ($t\in[0,\infty]$)
the part of $\varGamma$ where the slope does not exceed $t\myp
n_2/n_1$. Consider the set
\begin{equation}\label{eq:X_n}
\calX(t):=\{x\in \calX:\,x_2/x_1\le t\myp n_2/n_1\}, \qquad
t\in[0,\infty].
\end{equation}
According to the association
$\CP\ni\varGamma\leftrightarrow\nu\in\varPhi_0$ described in
Section~\ref{sec2.1}, for each $t\in[0,\infty]$ the polygonal line
$\varGamma(t)$ is determined by a truncated configuration
$\{\nu(x),\,x\in\calX(t)\}$, hence its right endpoint
$\xi(t)=(\xi_1(t),\xi_2(t))$ is given by
\begin{equation}\label{xi(t)}
\xi(t)=\sum_{x\in \calX(t)}x\myp\nu(x), \qquad t\in[0,\infty].
\end{equation}
In particular, $\calX(\infty)=\calX$, \,$\xi(\infty)=\xi$. Similarly
to (\ref{E_i'}),
\begin{equation}\label{eq:E(t)}
\EE_{z}[\xi(t)]= \sum_{k=1}^{\infty} k a_{k}\sum_{x\in
\calX(t)}x\mypp \rme^{-k\langle\alpha,x\rangle}, \qquad
t\in[0,\infty].
\end{equation}
Recall that the function $g^*(t)=(g_1^*(t),g_2^*(t))$
is defined in (\ref{eq:g*}).
\begin{theorem}\label{th:3.2}
For each\/ $t\in[0,\infty]$\textup{,}
\begin{equation}\label{sh}
\lim_{n\to\infty} n_j^{-1}\EE_{z}[\xi_j(t)]=g^*_j(t)\qquad (j=1,2).
\end{equation}
\end{theorem}
\begin{proof}
Theorem \ref{th:delta12} implies that (\ref{sh}) holds for
$t=\infty$. Assume that $t<\infty$ and let $j=1$ (the case $j=2$ is
considered in a similar manner). Setting for brevity $c_n:=n_2/n_1$
and arguing as in the proof of Theorem \ref{th:delta12} (see
(\ref{E_i'}), (\ref{E_1F}) and (\ref{E_1'})), from (\ref{eq:E(t)})
we obtain
\begin{align}
\EE_{z}[\xi_1(t)]& =\sum_{k,\myp m=1}^\infty k a_{k}\myp
m\myp\mu(m)\sum_{x_1=1}^\infty
x_1\rme^{-km\alpha_1x_1}\sum_{x_2=0}^{\hat x_2}
\rme^{-km\alpha_2x_2}
\notag\\
\label{Z1} &=\sum_{k,\myp m=1}^\infty k a_{k}\myp
m\myp\mu(m)\sum_{x_1=1}^\infty
x_1\rme^{-km\alpha_1x_1}\,\frac{1-\rme^{-km\alpha_2(\hat{x}_2+1)}}{1-\rme^{-km\alpha_2}},
\end{align}
where ${\hat{x}}_2={\hat{x}}_2(t)$ denotes the integer part of
$t\myp c_n\myp x_1$, so that
\begin{equation}\label{eq:x*}
0\le t\myp c_n\myp x_1-\hat{x}_2<1.
\end{equation}
Aiming to replace $\hat{x}_2+1$ by $t\myp c_n\myp x_1$ in
(\ref{Z1}), we recall (\ref{eq:alpha2:1}) and rewrite the sum over
$x_1$ as
\begin{equation}\label{eq:sum+Delta}
\sum_{x_1=1}^\infty
x_1\rme^{-km\alpha_1x_1}\bigl(1-\rme^{-km\alpha_1tx_1\myn}\bigr)
+\Delta_{k,\myp{}m}(t,\alpha),
\end{equation}
where
\begin{align*}
\Delta_{k,\myp{}m}(t,\alpha):= \sum_{x_1=1}^\infty
x_1\rme^{-km\alpha_1x_1(1+t)}\!
\left(1-\rme^{-km\alpha_2\myp(\hat{x}_2+1-t\myp c_n x_1)}\right).
\end{align*}
Using that $0<\hat{x}_2+1-t\myp c_n\myp x_1\le 1$ (see
(\ref{eq:x*})) and applying estimate (\ref{eq:exp_bound}), we
obtain, uniformly in $k,m\ge1$ and $t\in[0,\infty]$,
\begin{align*}
0<\frac{\Delta_{k,\myp{}m}(t,\alpha)}{1-\rme^{-km\alpha_2}}&\le
\sum_{x_1=1}^\infty x_1\rme^{-km\alpha_1x_1}=
\frac{\rme^{-km\alpha_1}}{(1-\rme^{-km\alpha_1\myn})^2}
=O(1)\,\frac{\rme^{-m\alpha_1/2}}{(km\alpha_1\myn)^{2}}\myp.
\end{align*}
Substituting this estimate into (\ref{Z1}) and using the condition
$A^+(2)<\infty$, we see that the error resulting from the
replacement of $\hat{x}_2+1$ by $t\myp c_n\myp x_1$ is dominated by
\begin{align*}
O\bigl(\alpha_1^{-2}\bigr)\sum_{k=1}^\infty \frac{|a_{k}|}{k}
\sum_{m=1}^\infty \frac{\rme^{-m\alpha_1/2}}{m}
=O\bigl(\alpha_1^{-2}\bigr)\myp
\ln\myn\bigl(1-\rme^{-\alpha_1/2}\bigr)
=O\bigl(\alpha_1^{-2}\ln\alpha_1\bigr).
\end{align*}
Returning to representation (\ref{Z1}) and evaluating the sum in
(\ref{eq:sum+Delta}), we find
\begin{equation}\label{UE_1}
\EE_{z}[\xi_1(t)] =\sum_{k,\myp m=1}^\infty k
a_{k}\left.\frac{m\myp\mu(m)}{1-\rme^{-km\alpha_2}}\cdot
\frac{\rme^{-km\alpha_1y}}{(1-\rme^{-km\alpha_1y})^2}\,
\right|_{\,y=1+t}^{\,y=1}+O\bigl(\alpha_1^{-2}\ln\alpha_1\bigr).
\end{equation}
Then, passing to the limit by Lebesgue's dominated convergence
theorem, similarly to the proof of Theorem \ref{th:delta12} (cf.\
(\ref{zeta32})) we get, as $n\to\infty$,
\begin{equation*}
n_1^{-1}\EE_{z}[\xi_1(t)]\to \frac{1}{\kappa^3}\sum_{k=1}^\infty
\frac{a_{k}}{k^{2}}\sum_{m=1}^\infty \frac{\mu(m)}{m^{2}}
\left(1-\frac{1}{(1+t)^2}\right) =\frac{t^2+2t}{(1+t)^2}\myp,
\end{equation*}
which coincides with $g^*_1(t)$, as claimed.
\end{proof}
There is a stronger version of Theorem~\ref{th:3.2}.
\begin{theorem}\label{th:8.1.1a}
Convergence in\/ \textup{(\ref{sh})} is uniform in\/
$t\in[0,\infty]$\textup{,} that is\textup{,}
\begin{equation*}
\lim_{n\to\infty}\sup_{0\le t\le\infty} \bigl|\myp
n_j^{-1}\EE_{z}[\xi_j(t)]-g^*_j(t)\bigr|=0\qquad (j=1,2).
\end{equation*}
\end{theorem}
For the proof, we shall use the following simple criterion of
uniform convergence proved in \cite[Lemma 4.3]{BZ4}.
\begin{lemma}\label{lm:8.1}
Let\/ $\{f_n(t)\}$ be a sequence of nondecreasing functions on a
finite interval\/ $[a,b\myp]$\textup{,} such that\textup{,} for
each\/ $t\in[a,b\myp]$\textup{,}
$\lim_{n\to\infty}f_n(t)=f(t)$\textup{,} where\/ $f(t)$ is a
continuous \textup{(}nondecreasing\textup{)} function on\/
$[a,b\myp]$. Then the convergence\/ $f_n(t)\to f(t)$ as\/
$n\to\infty$ is uniform on\/ $[a,b\myp]$.
\end{lemma}
\begin{proof}[Proof of Theorem\/ \textup{\ref{th:8.1.1a}}] Suppose that
$j=1$ (the case $j=2$ is handled similarly). Note that for each $n$
the function
\begin{equation*}
f_n(t):=n_1^{-1}\EE_{z}[\myp\xi_1(t)]=\frac{1}{n_1} \sum_{x\in
\calX(t)} x_1\myp \EE_{z}[\nu(x)]
\end{equation*}
is nondecreasing in $t$. Therefore, by Lemma \ref{lm:8.1} the
convergence in (\ref{sh}) is uniform on any interval $[0,t^*]$
\,($t^*<\infty$). Since $n_1^{-1}\EE_{z}[\xi_1(\infty)]\to
g^*_1(\infty)$ and the function $g^*_1(t)$ is continuous at infinity
(see (\ref{eq:g*})), it remains to show that for any $\varepsilon>0$
there is $t^*$ such that, for all large enough $n_1,\,n_2$ and all
$t\ge t^*$,
\begin{equation}\label{eq:E<epsilon}
n_1^{-1}\EE_{z}|\myp\xi_1(\infty)-\xi_1(t)|\le \varepsilon.
\end{equation}
To this end, on account of (\ref{UE_1}) we have
\begin{equation}\label{eq:Ez}
\EE_{z}[\xi_1(\infty)-\xi_1(t)] =\sum_{k,\myp m=1}^\infty k
a_{k}\myp \frac{m\myp\mu(m)}{1-\rme^{-km\alpha_2}}\cdot
\frac{\rme^{-km\alpha_1(1+t)}}{\bigl(1-\rme^{-km\alpha_1(1+t)}\bigr)^2}+O\bigl(\alpha_1^{-2}\ln\alpha_1\bigr).
\end{equation}
Note that by inequality (\ref{eq:exp_bound}), uniformly in
$k,m\ge1$,
\begin{equation*}
\frac{\rme^{-km\alpha_2}}{1-\rme^{-km\alpha_2}}\cdot
\frac{\rme^{-km\alpha_1(1+t)}}{(1-\rme^{-km\alpha_1(1+t)})^2}=
\frac{O(1)}{\alpha_1^{2}\alpha_2 (km)^{3}(1+t)^{2}}\myp.
\end{equation*}
Returning to (\ref{eq:Ez}) and using the condition $A^+(2)<\infty$,
we obtain, uniformly in $t\ge t^*$,
\begin{align*}
\alpha_1^{2}\alpha_2\EE_{z}[\xi_1(\infty)-\xi_1(t)]
&=\frac{O(1)}{(1+t)^{2}} \sum_{k=1}^\infty
\frac{|a_{k}|}{k^{2}}\sum_{m=1}^\infty
\frac{1}{m^{2}}=\frac{O(1)}{(1+t^*)^{2}}\myp,
\end{align*}
whence by (\ref{alpha}) we get (\ref{eq:E<epsilon}).
\end{proof}
\section{Refined asymptotics of the expectation}\label{sec4}
We need to sharpen the asymptotic estimate $\EE_z(\xi)-n=o(|n|)$
provided by Theorem \textup{\ref{th:delta12}}.
\begin{theorem}\label{th:4.1}
Under the condition $A^+(1)<\infty$\textup{,} we have
$\EE_{z}(\xi)-n=O(|n|^{2/3})$ as $n\to\infty$.
\end{theorem}
For the proof of Theorem \ref{th:4.1}, some preparations are
required.
\subsection{Integral approximation of sums}\label{sec4.1}
Let a function $f\colon \RR^2_+\to\RR$ be continuous and absolutely
integrable on $\RR^2_+$\myp, together with its partial derivatives
up to the second order. Set
\begin{equation}\label{F0}
F(h):=\sum_{x\in\ZZ^2_+}f(hx), \qquad h>0
\end{equation}
(as one can verify, the above conditions on $f$ ensure that the
series in (\ref{F0}) is absolutely convergent \cite[p.\,21]{BZ4}),
and assume that for some $\beta>2$
\begin{equation}\label{beta}
F(h)=O\bigl(h^{-\beta}\bigr),\qquad h\to\infty.
\end{equation}
Consider the Mellin transform of $F(h)$ (see, e.g., \cite[Ch.\,VI,
\S\myp9]{Widder}),
\begin{equation}\label{Mel}
\widehat{F}(s):=\int_0^\infty h^{s-1}F(h)\,\dif{}h\myp,
\end{equation}
and set
\begin{equation}\label{Delta}
\Delta_f(h):=F(h)-\frac{1}{h^2}\int_{\RR^2_+} f(x)\,\dif{x}, \qquad
h>0.
\end{equation}
The following general lemma can be proved using the well-known
Euler--Maclaurin summation formula (see details in \cite[Lemmas 5.1
and 5.3, pp.\,20--22]{BZ4}).
\begin{lemma}\label{lm:Delta1}
The function\/ $\widehat{F}(s)$ is meromorphic in the strip\/ $1<\Re
s<\beta$\textup{,} with a single\/
\textup{(}\myn{}simple\myp\textup{)} pole at\/ $s=2$.
Moreover\textup{,} $\widehat{F}(s)$ satisfies the identities
\begin{align}\label{Muntz2}
\widehat{F}(s)&=\int_0^\infty h^{s-1}\Delta_f(h)\,\dif{}h\myp,\qquad
1<\Re s<2,\\
\label{eq:inverse}
\Delta_f(h)&=\frac{1}{2\pi
i}\int_{c-i\infty}^{c+i\infty} h^{-s} \widehat{F}(s)\,\dif{}s,\qquad
1<c<2.
\end{align}
\end{lemma}
\subsection{Proof of Theorem\/ \textup{\ref{th:4.1}}}\label{sec4.2}
Our argumentation follows the same lines as in a similar result in
\cite[pp.\,22--27]{BZ4} for distributions determined by coefficients
(\ref{eq:b-k-r}) (with $\rho=1$). For the reader's convenience, we
repeat all the steps but skip some word-by-word repetitions, giving
specific references to \cite{BZ4}.
Let us consider $\xi_1$ (for $\xi_2$ the proof is similar).
Recalling the notations $f(x)$ and $F(h)$ introduced in Section
\ref{sec3.2} (see (\ref{f+}) and (\ref{F}), respectively), we have,
according to (\ref{E_1}),
\begin{equation}\label{E_z''}
\EE_{z}(\xi_1)=\sum_{k,\myp m=1}^\infty a_{k}\myp \mu(m) F(km),
\end{equation}
where
\begin{gather*}
F(h)=\sum_{x\in\ZZ^2_+} f(hx)=\frac{h\mypp
\rme^{-\alpha_1h}}{(1-\rme^{-\alpha_1h})^2\myp(1-\rme^{-\alpha_2h})}\myp,\qquad
h>0,\\
f(x)=x_1\rme^{-\langle\alpha,x\rangle},\qquad x\in \RR^2_+\myp.
\end{gather*}
Note that
\begin{equation*}
\int_{\RR^2_+}f(x)\,\dif{}x=\int_0^\infty
x_1\rme^{-\alpha_1x_1}\,\dif{}x_1\int_0^\infty
\rme^{-\alpha_2x_2}\,\dif{}x_2= \frac{1}{\alpha_1^2\myp
\alpha_2}\myp.
\end{equation*}
Moreover, using
(\ref{eq:kappa}) and
(\ref{eq:alpha2:1}) we have (cf.\ (\ref{zeta32}))
\begin{equation}\label{eq:int(f)} \sum_{k,\myp
m=1}^\infty \frac{a_{k}\myp \mu(m)}{(km)^2\alpha_1^2\alpha_2}=
\frac{n_1}{\kappa^3}\sum_{k=1}^\infty
\frac{a_{k}}{k^2}\sum_{m=1}^\infty \frac{\mu(m)}{m^2}\equiv n_1.
\end{equation}
Subtracting (\ref{eq:int(f)}) from (\ref{E_z''}), we obtain the
representation
\begin{equation}\label{dif1}
\EE_{z}(\xi_1)-n_1=\sum_{k,\myp m=1}^\infty a_{k}\myp\mu(m)\,
\Delta_{f}\mynn(km),
\end{equation}
where $\Delta_f(h)$ is defined in (\ref{Delta}). Clearly, the
functions $f$ and $F$ satisfy the hypotheses of Lemma
\ref{lm:Delta1} (with $\beta=\infty$). Setting $c_n:=n_2/n_1$ and
using (\ref{eq:alpha2:1}), the Mellin transform of $F(h)$ defined by
(\ref{Mel}) can be represented as
\begin{equation}\label{eq:M_alpha}
\widehat{F}(s)=\alpha_1^{-s-1}\widetilde F(s),
\end{equation}
where
\begin{equation}\label{tildeM}
\widetilde{F}(s):=\int_0^\infty \frac{y^s\myp
\rme^{-y}}{(1-\rme^{-y})^2\, (1-\rme^{-y/c_n})}\,\dif{}y,\qquad \Re
s>2\myp.
\end{equation}
It is easy to verify (see \cite[p.~23]{BZ4} for details) that the
analytic continuation of expression (\ref{tildeM}) into domain
$1<\Re s<2$\, is explicitly given by
\begin{equation}\label{tildeM2}
\widetilde{F}(s)=J(s)+
c_n\myp\zeta(s-1)\mypp\Gamma(s)+\frac{1}{2}\,\zeta(s)\mypp\Gamma(s+1),
\end{equation}
where $\Gamma(s)= \int_0^\infty u^{s-1}\,\rme^{-u}\,\dif{u}$ is the
gamma function, $\zeta(s)=\sum_{k=1}^\infty k^{-s}$ is the Riemann
zeta function and
\begin{equation}\label{J}
J(s):=\int_0^\infty \frac{y^s\myp \rme^{-y}}
{(1-\rme^{-y})^2}\left(\frac{1}{1-\rme^{-y/c_n}}-\frac{c_n}{y}-\frac12\right)
\dif{y}.
\end{equation}
Note that for $\Re s>0$ the integral in (\ref{J}) is absolutely
convergent and therefore
$J(s)$
is regular. Furthermore, it is well known that $\Gamma(s)$ is
analytic for $\Re s > 0$ \cite[\S\mypp{}4.41, p.\,148]{Titch2},
whereas $\zeta(s)$ has a single pole at point $s = 1$
\cite[\S\mypp{}4.43, p.\,152]{Titch2}. Thus, the right-hand side of
(\ref{tildeM2}) is meromorphic in the half-plane $\Re s>0$, with
simple poles at $s=1$ and $s=2$\myp.
Using (\ref{eq:inverse}) and (\ref{eq:M_alpha}), and recalling
formulas (\ref{eq:A2-}) and (\ref{eq:zeta^{-1}}), we can rewrite
(\ref{dif1}) as
\begin{align}
\notag \EE_{z}(\xi_1)-n_1&=\frac{1}{2\pi\myp \rmi} \sum_{k,\myp
m=1}^\infty a_{k}\myp \mu(m) \int_{c-\rmi\infty}^{c+\rmi\myp\infty}
\frac{\widetilde{F}(s)}{\alpha_1^{s+1}(km)^{s}}\,\dif{}s\\
\label{dif4} &=\frac{1}{2\pi
\rmi}\int_{c-\rmi\myp\infty}^{c+\rmi\myp\infty}
\frac{A(s)}{\alpha_1^{s+1}\zeta(s)}\,\widetilde{F}(s)\,\dif{}s
\qquad(1< c< 2).
\end{align}
Noting that $\zeta(s)\ne 0$ for $\Re s \ge 1$, let us show that the
integration contour $\Re s=c$ in (\ref{dif4}) can be moved to $\Re
s=1$. By the Cauchy theorem, it suffices to check that
\begin{equation}\label{iT->0}
\lim_{t\to\infty}\int_{1-\rmi\myp t}^{c+\rmi\myp
t}\frac{A(s)}{\alpha_1^{s+1}\zeta(s)}\,\widetilde{F}(s)\,\dif{}s=0,\qquad(1<
c<2).
\end{equation}
To this end, note that for $s=\sigma+\rmi\myp t$ with $1\le\sigma\le
c<2$
\begin{equation}\label{|A|<}
|A(s)|\le A^+(1)<\infty,\qquad \bigl|\alpha_1^{-s-1}\bigr| \le
\alpha_1^{-c-1}\myp,
\end{equation}
whereas integration by parts in (\ref{J}) yields a uniform estimate
\begin{equation}\label{J(it)}
J(s)= O\bigl(t^{-2}\bigr),\qquad t\to\infty.
\end{equation}
Furthermore, we have the following asymptotic estimates as
$t\to\infty$, uniform in the strip $\Re s\in[1,c]$ (see
\cite[\S\mypp{}4.42, p.\,151]{Titch2} for (\ref{GFE}),
\cite[Theorem~1.9, p.\:25]{Iv} for (\ref{zeta_123}) and
\cite[Eq.\;(3.11.8), p.\;60]{Titch1} for (\ref{zeta_1}))
\begin{gather}
\label{GFE} \Gamma(s)=O\myp(|t|^{\sigma-1/2}\mypp
\rme^{-\pi|t|/2}),\qquad \Gamma(s+1)=O\myp(|t|^{\sigma+1/2}\mypp
\rme^{-\pi|t|/2}),\\
\label{zeta_123} \zeta(s)=O\myp(\ln\myn|t|),\qquad
\zeta(s-1)=O\myp(t^{1-\sigma/2}\ln\myn|t|),\\[.3pc]
\label{zeta_1} \zeta(s)^{-1}=O(\ln\myn|t|).
\end{gather}
Substituting estimates (\ref{GFE}) and (\ref{zeta_123}) into
(\ref{tildeM2}), we get $\widetilde{F}(s)=O(t^{-2})$, and on account
of (\ref{|A|<}) and (\ref{J(it)}) we see that (\ref{iT->0}) follows.
Hence, representation (\ref{dif4}) takes the form
\begin{equation*}
\EE_{z}(\xi_1)-n_1=\frac{1}{2\pi}
\int_{-\infty}^\infty \frac{A(1+\rmi\myp
t)}{
\alpha_1^{2+\rmi t}
\mypp\zeta(1+\rmi\myp t)}\,\widetilde{F}(1+\rmi\myp
t)\,\dif{}t = O(\alpha_1^{-2})=O(|n|^{2/3}),
\end{equation*}
according to (\ref{alpha}). The proof of Theorem \ref{th:4.1} is
complete.
\begin{remark}
If condition $A^+(\sigma)<\infty$ is satisfied with some
$\sigma\in(0,1)$, then the statement of Theorem \ref{th:4.1} can be
enhanced to $\EE_{z}(\xi)-n=o(|n|^{2/3})$ (cf.\ \cite[p.~26]{BZ4}).
This is the case for all examples in Section \ref{sec2.3}.
\end{remark}
\section{Asymptotics of higher-order moments}\label{sec5}
In this section, we again assume that $A^+(2)<\infty$.
\subsection{Second-order moments}\label{sec5.1}
As before, denote $a_z:=\EE_{z}(\xi)$, and let $K_z:=
\Cov(\xi,\xi)= \EE_{z}
(\xi-a_z\myn)^{\myn\topp}(\xi-a_z\myn)$
be the covariance matrix of the random vector
\,$\xi=\sum_{x\in\calX} x\myp\nu(x)$. Recalling that the random
variables $\nu(x)$ are independent for different $x\in \calX$ and
using (\ref{eq:nu2}), we see that the elements
$K_z(i,j)=\Cov(\xi_i,\xi_j)$ ($i,j\in\{1,2\}$) of the matrix $K_z$
are given by
\begin{equation}\label{D_zxi}
K_z(i,j)=\sum_{x\in \calX}x_i x_j \Var[\nu(x)]=\sum_{x\in \calX} x_i
x_j \sum_{k=1}^\infty k^2 a_{k}\mypp z^{kx}.
\end{equation}
\begin{theorem}\label{th:K}
As\/ $n\to\infty$\textup{,}
\begin{equation}\label{eq:Sigma}
K_z(i,j)\sim B_{ij}\,(n_1n_2)^{2/3},\qquad i,j\in\{1,2\},
\end{equation}
where the matrix $B:=(B_{ij})$ is given by
\begin{equation}\label{eq:Sigma1}
B= \kappa^{-1}\left(\arraycolsep=.20pc\begin{array}{cc}
2\myp n_1/n_2&1\\[.3pc]
1&2\myp n_2/n_1\myn
\end{array}
\right)\!.
\end{equation}
\end{theorem}
\begin{proof}
Let us consider $K_z(1,1)$ (the other elements of $K_z$ are analyzed
in a similar manner). Substituting (\ref{alpha}) into (\ref{D_zxi}),
we obtain
\begin{equation}\label{D_zxi+}
K_z(1,1)=\sum_{x\in \calX} x_1^2\sum_{k=1}^\infty k^2
a_{k}\myp\rme^{-k\langle\alpha,x\rangle}.
\end{equation}
Using the M\"obius inversion formula (\ref{eq:Mobius}), similarly to
(\ref{E_1'}) the double sum in (\ref{D_zxi+}) can be rewritten in
the form
\begin{align}
\notag K_z(1,1)&=\sum_{m=1}^\infty m^2\mu(m) \sum_{k=1}^\infty k^2
a_{k}\sum_{x\in\ZZ^2_+}
x_1^2\myp\rme^{-km\langle\alpha,x\rangle}\\
\notag &=\sum_{k,\myp m=1}^\infty m^2\mu(m)\mypp k^2 a_{k}
\sum_{x_1=1}^\infty x_1^2\myp\rme^{-km\alpha_1 x_1}
\sum_{x_2=0}^{\infty}
\rme^{-km\alpha_2 x_2}\\
\label{D1_1(t)} &=\sum_{k,\myp m=1}^\infty m^2\mu(m)\mypp k^2
a_{k}\, \frac{\,\rme^{-km\alpha_1}
(1+\rme^{-km\alpha_1\myn})}{(1-\rme^{-k
m\alpha_1\myn})^3(1-\rme^{-km\alpha_2})} \myp.
\end{align}
By estimate (\ref{eq:exp_bound}), the general term in series
(\ref{D1_1(t)}) is bounded by $\alpha_1^{-3}\alpha_2^{-1}\,
O(|a_{k}|\myp k^{-2}\myp m^{-2})$, uniformly in $k,m$, and
furthermore (see (\ref{eq:A2-}))
\begin{equation*}
\sum_{k,\myp m=1}^\infty \frac{|a_{k}|}{k^2 m^2}= A^+(2)\,\zeta(2)
<\infty.
\end{equation*}
Therefore, Lebesgue's dominated convergence theorem yields
\begin{equation}\label{eq:->A4}
\alpha_1^3\alpha_2\myp K_z(1,1) \to 2\sum_{k=1}^\infty
\frac{a_{k}}{k^2}\sum_{m=1}^\infty \frac{\mu(m)}
{m^2}=\frac{2A(2)}{\zeta(2)},\qquad \alpha_1,\alpha_2\to0.
\end{equation}
Hence, using (\ref{alpha}), (\ref{delta12}) and (\ref{eq:kappa}),
from (\ref{eq:->A4}) we get, as $n\to\infty$,
\begin{equation*}
K_z(1,1)\sim \frac{2\myp A(2)}{\zeta(2) \mypp \kappa^4}
\,(n_1/n_2)\mypp(n_1n_2)^{2/3}=B_{11}\myp(n_1n_2)^{2/3},
\end{equation*}
as required (cf.\ (\ref{eq:Sigma}), (\ref{eq:Sigma1})).
\end{proof}
The next lemma is a direct corollary of Theorem \ref{th:K}.
\begin{lemma}\label{lm:detK} As\/
$n\to\infty$\textup{,}
\begin{equation*}
\det K_z\sim 3\myp \kappa^{-2}(n_1n_2)^{4/3}.
\end{equation*}
\end{lemma}
Lemma \ref{lm:detK} implies that the matrix $K_z$ is non-degenerate,
at least asymptotically as $n\to\infty$. In fact, from (\ref{D_zxi})
it is easy to see (e.g., using the Cauchy--Schwarz inequality
together with the characterization of the equality case) that
$K_z$ is positive definite; in particular, $\det K_z>0$ and
hence $K_z$ is invertible. Let $V_z=K_z^{-1/2}$ be the (unique)
square root of the matrix $K_z^{-1}$,
that is, a symmetric, positive definite matrix such that
$V_z^2=K_z^{-1}$.
Recall that the matrix norm induced by the Euclidean vector norm
$|\,{\cdot}\,|$ is defined by $\|A\|:=\sup_{|x|=1}|x A|$.
We need some general facts about this matrix norm (see
\cite[\S\myp7.2, pp.\,33--34]{BZ4} for simple proofs and
bibliographic comments).
\begin{lemma}
\label{lm:7.1.1} If\/ $A$ is a real matrix then $\|A^{\topp}\mynn
A\|=\|A\|^2$.
\end{lemma}
\begin{lemma}
\label{lm:|A|} If\/ $A=(a_{ij})$ is a\/ real $d\times d$
matrix\myp\textup{,} then\/
\begin{equation}\label{eq:m-norm}
\frac{1}{d}\sum_{i,\myp{}j=1}^d a_{ij}^2\le
\|A\|^2\le\sum_{i,\myp{}j=1}^d a_{ij}^2\myp.
\end{equation}
\end{lemma}
\begin{lemma}
\label{lm:7.1.2} Let\/ $A$ be a symmetric\/ $2\times 2$ matrix
with\/ $\det A\ne 0$. Then
\begin{equation}\label{eq:d=2}
\|A^{-1}\|=\frac{\|A\|}{|\mynn\det A|}\myp.
\end{equation}
\end{lemma}
Let us now estimate the norms of the matrices $K_z$ and
$V_z=K_z^{-1/2}$.
\begin{lemma}\label{lm:K_z}
As\/ $n\to\infty$\textup{,} one has\/ $\|K_z\|\asymp |n|^{4/3}$.
\end{lemma}
\begin{proof}
Lemma \ref{lm:|A|} and Theorem \ref{th:K} imply
\begin{equation*}
\|K_z\|^2\asymp \sum_{i,\myp{}j=1}^2 K_z(i,j)^2\asymp
(n_1n_2)^{4/3}\asymp |n|^{8/3}\qquad (n\to\infty),
\end{equation*}
and the required estimate follows.
\end{proof}
\begin{lemma}\label{NV}
For the matrix\/ $V_z=K_z^{-1/2}$\textup{,} one has\/ $\|V_z\|\asymp
|n|^{-2/3}$ as\/ $n\to\infty$.
\end{lemma}
\begin{proof}
Using Lemmas \ref{lm:7.1.1} and \ref{lm:7.1.2} we have
\begin{equation*}
\|V_z\|^2=\|V_z^2\|=\|K_z^{-1}\|=\frac{\|K_z\|}{\det K_z}\myp,
\end{equation*}
and an application of Lemmas \ref{lm:detK} and \ref{lm:K_z}
completes the proof.
\end{proof}
\subsection{Auxiliary estimates}\label{sec5.2}
Denote
\begin{equation}\label{eq:nu0}
\nu_0(x):=\nu(x)-\EE_{z}[\nu(x)],\qquad x\in \calX,
\end{equation}
and consider the moments of order $q\in\NN$
\begin{equation}\label{eq:m-mu}
m_q(x):=\EE_{z}\myn\bigl[\nu(x)^q\bigr],\qquad
\mu_q(x):=\EE_{z}\myn\bigl|\nu_0(x)^q\bigr|
\end{equation}
(for simplicity, we suppress the dependence on $z$).
Let us note a simple general inequality (cf.\
\mbox{\cite[Lemma~6.2]{BZ4}}).
\begin{lemma}\label{lm:mu<m}
For each\/
$q\ge1$ and all\/ $x\in \calX$\textup{,}
\begin{equation}\label{eq:mu<m}
\mu_q(x)\le 2^q \myp m_q(x).
\end{equation}
\end{lemma}
\proof Using the elementary inequality $(a+b)^q\le 2^{q-1}(a^{\myp
q}+b^{\myp q})$ for any $a,b>0$ and $q\ge 1$ (which follows from
H\"older's inequality for the function $y=x^q$), we obtain
\begin{align*}
\mu_q(x)&\le
\EE_{z}\bigl[\bigl(\nu(x)+m_1(x)\bigr)^q\bigr]\le2^{q-1}\bigl(m_q(x)+m_1(x)^q\bigr)\le
2^q \myp m_q(x),
\end{align*}
where
we used Lyapunov's inequality $m_1(x)^q\le
m_q(x)$.
\endproof
The following two lemmas are useful for estimation of higher-order
moment sums.
\begin{lemma}\label{lm:6.2}
For\/ $q\in\NN$\textup{,} the function
\begin{equation}\label{eq:q}
S_q(\theta):=\sum_{x=1}^\infty x^{q-1}\rme^{-\theta x},\qquad
\theta>0,
\end{equation}
admits a representation
\begin{equation}\label{eq:S}
S_q(\theta)=\sum_{j=1}^{q} c_{j,\myp q}\,\frac{\rme^{-\theta
j}}{(1-\rme^{-\theta})^j},\qquad \theta>0,
\end{equation}
with some constants\/ $c_{j,\myp q}>0$
\,\textup{(}$j=1,\dots,q$\textup{)}\myp\textup{;} in
particular\textup{,} $c_{q,\myp q}=(q-1)!$\mypp.
\end{lemma}
\proof In the case $q=1$, expression (\ref{eq:q}) is reduced to a
geometric series
\begin{align*}
S_1(\theta)=\sum_{x=1}^\infty \rme^{-\theta x}=\frac{\rme^{-\theta
}}{1-\rme^{-\theta}}\myp,
\end{align*}
which is a particular case of (\ref{eq:S}) with $c_{1,1}:=1$. Assume
now that (\ref{eq:S}) is valid for some $q\ge1$. Then,
differentiating identities (\ref{eq:q}) and (\ref{eq:S}) with
respect to $\theta$, we obtain
\begin{align*}
S_{q+1}(\theta)=-\frac{\dif}{\dif \theta}
S_q(\theta)&=\sum_{j=1}^{q} c_{j,\myp q}\left(\frac{j\mypp
\rme^{-\theta j}}{(1-\rme^{-\theta})^j}+ \frac{j\mypp\rme^{-\theta
(j+1)}}{(1-\rme^{-\theta})^{j+1}}\right)\\
&=\sum_{j=1}^{q+1} c_{j,\myp q+1}\frac{\rme^{-\theta
j}}{(1-\rme^{-\theta})^j}\myp,
\end{align*}
where we set
\begin{equation*}
c_{j,\myp{}q+1}:=\left\{\begin{array}{ll}
c_{1,\myp{}q}\myp,&j=1,\\[.2pc]
j\myp c_{j,\myp{}q}+(j-1)\mypp c_{j-1,\myp{}q}\myp,\ \ &2\le j\le q,\\[.2pc]
q\myp c_{q,\myp{}q}\myp,& j=q+1.
\end{array}
\right.
\end{equation*}
In particular, $c_{q+1,\myp q+1}=q\myp
c_{q,\myp{}q}=q\myp(q-1)!=q!$. Thus, formula (\ref{eq:S}) holds for
$q+1$ and hence, by induction, for all $q\ge1$.
\endproof
\begin{lemma}\label{lm:C_k}
For each $q\in\NN$, there exists a positive constant
$C_q$ such that, for all $\theta>0$,
\begin{equation}\label{eq:Cc}
0< S_q(\theta)\le
\frac{C_q\mypp\rme^{-\theta}}{(1-\rme^{-\theta})^q}\myp.
\end{equation}
\end{lemma}
\proof Observe that for $j=1,\dots,q$ and all $\theta>0$
\begin{equation*}
\frac{\rme^{-\theta
j}}{(1-\rme^{-\theta})^j}\le\frac{\rme^{-\theta}}{(1-\rme^{-\theta})^q}\myp.
\end{equation*}
Substituting these inequalities into (\ref{eq:S}) and recalling that
the coefficients $c_{j,\myp q}$ are positive, we obtain
(\ref{eq:Cc}) with
$C_q:=\sum_{j=1}^q
c_{j,\myp q}$.
\endproof
\subsection{Asymptotics of moment sums}\label{sec5.3}
According to (\ref{eq:cumulants-xi}) and (\ref{alpha}), the
cumulants of $\xi_j=\sum_{x\in\mathcal{X}}x_j\myp\nu(x)$ ($j=1,2$)
are given by
\begin{equation}\label{varkappa_zxi+}
\varkappa_q[\xi_j]=\sum_{x\in \calX}
x_j^q\myp\varkappa_q(x)=\sum_{x\in \calX} x_j^q\sum_{k=1}^\infty k^q
a_{k}\mypp \rme^{-k\langle \alpha,x\rangle},\qquad q\in\NN.
\end{equation}
\begin{lemma}\label{lm:kappa_q}
For each $q\in\NN$ and $j=1,2$\textup{,}
\begin{equation}\label{eq:kappa_q}
\varkappa_q[\xi_j]\asymp |n|^{(q+2)/3},\qquad n\to\infty.
\end{equation}
\end{lemma}
\proof Let $j=1$ (the case $j=2$ is treated in a similar fashion).
Using the M\"obius inversion formula (\ref{eq:Mobius}), similarly to
(\ref{E_1'}) the right-hand side of (\ref{varkappa_zxi+}) (with
$j=1$) can be rewritten as
\begin{align}
\notag \varkappa_q[\xi_1]&=\sum_{k,\myp m=1}^\infty m^q\mu(m) \mypp
k^q a_{k}\sum_{x\in\ZZ^2_+}
x_1^q\myp\rme^{-km\langle\alpha,x\rangle}\\
\notag &=\sum_{k,\myp m=1}^\infty m^q\mu(m)\mypp k^q a_{k}
\sum_{x_1=1}^\infty x_1^q\myp\rme^{-km\alpha_1 x_1}
\sum_{x_2=0}^{\infty}
\rme^{-km\alpha_2 x_2}\\
\label{D1_1(t)'} &=\sum_{k,\myp m=1}^\infty m^q\mu(m)\mypp k^q
a_{k}\,S_{q+1}(km\alpha_1)\,(1-\rme^{-km\alpha_2})^{-1},
\end{align}
where in the last line we used notation (\ref{eq:q}). Lemma
\ref{lm:C_k} and inequality (\ref{eq:exp_bound}) show that the
general term in series (\ref{D1_1(t)'}) is bounded in absolute
value, uniformly in $k$ and $m$, as follows
$$
m^q \mypp k^q
|a_{k}|\,\frac{C_{q+1}\mypp\rme^{-km\alpha_1}}{(1-\rme^{-km\alpha_1})^{q+1}\myp(1-\rme^{-km\alpha_2})}=
\frac{O(1)\mypp |a_k|}{k^2m^2\alpha_1^{q+1}\alpha_2}
$$
(cf.\ (\ref{eq:2,1})). Hence, expanding $S_{q+1}(km\alpha_1)$ by
Lemma \ref{lm:6.2}, we can pass to the limit in (\ref{D1_1(t)'}) as
$\alpha_1,\alpha_2\to0$ to obtain
\begin{align}
\notag
\alpha_1^{q+1}\alpha_2\mypp\varkappa_q[\xi_1]
&=\sum_{j=1}^{q+1}c_{j,\myp q+1}\sum_{k,\myp m=1}^\infty
m^q\mu(m)\mypp k^q a_{k}\,\frac{\alpha_1^{q+1}\alpha_2\mypp
\mypp\rme^{-kmj\myp\alpha_1}}{(1-\rme^{-km\alpha_1})^{j}\myp(1-\rme^{-km\alpha_2})}\\
\label{eq:q!kappa} &\to c_{q+1,\myp q+1}\sum_{k,\myp m=1}^\infty
\frac{\mu(m)\mypp a_{k}}{k^2
m^2}=q!\,\frac{A(2)}{\zeta(2)}=q!\,\kappa^3\myp.
\end{align}
Finally, according to (\ref{alpha}) we have
$\alpha_1^{q+1}\alpha_2\asymp |n|^{-(q+2)/3}$, and hence
(\ref{eq:q!kappa}) implies (\ref{eq:kappa_q}).
\endproof
There is a similar upper asymptotic bound for the mixed cumulants.
\begin{lemma}\label{lm:cum} For each $q\in \NN$ and any\/
$t_1,t_2\in\RR$\textup{,}
$$
\varkappa_q[t_1\xi_1+t_2\xi_2]=O\bigl(|n|^{(q+2)/3}\bigr),\qquad
n\to\infty.
$$
\end{lemma}
\proof Similarly to representation (\ref{varkappa_zxi+}), we have
$$
\varkappa_q[t_1\xi_1+t_2\xi_2]=\sum_{x\in\calX}(t_1 x_1+t_2
x_2)^q\myp\varkappa_q(x).
$$
Hence, by the inequality $(a+b)^q\le 2^{q-1}(a^{\myp q}+b^{\myp q})$
(already used in the proof of Lemma \ref{lm:mu<m}), we obtain
\begin{align}
\notag
\bigl|\varkappa_q[t_1\xi_1+t_2\xi_2]\bigr|&\le
\sum_{x\in\calX}|t_1
x_1+t_2 x_2|^q\myp|\varkappa_q(x)|\\
\label{eq:<|a|} &\le \sum_{x\in\calX}\bigl(|t_1|^q x_1^q+|t_2|^q
x_2^q\bigr)\sum_{k=1}^\infty
k^q|a_k|\mypp\rme^{-k\langle\alpha,x\rangle}.
\end{align}
Repeating the arguments used in the proof of Lemma \ref{lm:kappa_q},
we see that the right-hand side of (\ref{eq:<|a|}) admits an
asymptotic bound $O(|n|^{(q+2)/3})$, and the lemma is proved.
\endproof
In view of relation (\ref{eq:moments-xi}),
Lemma \ref{lm:kappa_q}
immediately yields
the following corollary.
\begin{lemma}\label{lm:6.6}
For each $q\in\NN$ and $j=1,2$\textup{,}
\begin{equation*}
\EE_z(\xi_j^q)\asymp |n|^{(q+2)/3},\qquad n\to\infty.
\end{equation*}
\end{lemma}
We also have a similar upper estimate for the centered moments.
\begin{lemma
\label{lm:6.7} For each\/ $q\in\NN$ and $j=1,2$\textup{,}
\begin{equation*}
\EE_{z}|\xi_j-\EE_{z}(\xi_j)|^{q} =O\bigl(|n|^{(q+2)/3}\bigr),\qquad
n\to\infty.
\end{equation*}
\end{lemma}
\proof Applying an inequality similar to (\ref{eq:mu<m}), we obtain
\begin{align*}
\EE_{z}|\xi_j-\EE_{z}(\xi_j)|^{q}&\le 2^{q-1}\EE_{z}(\xi_j^q)\asymp
|n|^{(q+2)/3},
\end{align*}
according to Lemma \ref{lm:6.6}.
\endproof
\begin{lemma}\label{lm:6.3}
For each\/ $q\in\NN$\textup{,}
\begin{equation}\label{eq:sum_m}
\sum_{x\in\calX} |x|^q\mypp m_q(x)=O\bigl(|n|^{(q+2)/3}\bigr),
\qquad n\to\infty.
\end{equation}
\end{lemma}
\begin{proof}
Using the elementary inequalities $
|x|^q\le (x_1+x_2)^q\le 2^{q-1}(x_1^q+x_2^q)$ and recalling
definition (\ref{eq:m-mu}), observe that
\begin{equation*}
\sum_{x\in\calX} |x|^q\mypp m_q(x)\le 2^{q-1}
\bigl(\EE_z(\xi_1^q)+\EE_z(\xi_2^q)\bigr),
\end{equation*}
whence (\ref{eq:sum_m}) readily follows by Lemma \ref{lm:6.6}.
\end{proof}
\begin{lemma}\label{lm:muk} As\/ $n\to\infty$\textup{,}
\begin{equation*}
\sum_{x\in \calX}|x|^3\mu_3(x)\asymp |n|^{5/3},
\end{equation*}
where $\mu_3(x):=\EE_{z}|\nu_0(x)^3|$ \,\textup{(}see
\textup{(\ref{eq:m-mu})}\textup{)}.
\end{lemma}
\begin{proof}
An upper bound $O(|n|^{5/3})$ follows from inequality
(\ref{eq:mu<m}) and Lemma \ref{lm:6.3}. On the other hand, we have
\begin{equation*}
\mu_3(x)=\EE_z|\nu(x)-m_1(x)|^3
\ge\EE_z(\nu(x)-m_1(x))^3=\varkappa_3(x),
\end{equation*}
using that the third-order centered moment coincides with the
third-order cumulant. Hence, on account of formula
(\ref{varkappa_zxi+}),
\begin{equation*}
\sum_{x\in \calX}|x|^3\mu_3(x)\ge \sum_{x\in \calX}
x_1^3\myp\varkappa_3(x)=\varkappa_3[\xi_1]\asymp |n|^{5/3},
\end{equation*}
according to Lemma \ref{lm:kappa_q} (with $q=3$).
\end{proof}
Let us introduce the \emph{Lyapunov coefficient}
\begin{equation}\label{L3}
L_z:=\|V_z\|^{3} \sum_{x\in \calX}|x|^3\mu_3(x).
\end{equation}
The next asymptotic estimate is an immediate consequence of Lemmas
\ref{NV} and \ref{lm:muk}.
\begin{lemma}\label{lm:7.1}
As\/ $n\to\infty$\textup{,} one has\/ $L_z\asymp |n|^{-1/3}$.
\end{lemma}
\section{Local limit theorem}\label{sec7}
The role of a local limit theorem in our approach is to yield the
asymptotics of the probability $\QQ_z\{\xi=n\}\equiv\QQ_z(\CP_n)$
appearing in the representation of the measure $\PP_n$ as a
conditional distribution, $\PP_n(\cdot)=\QQ_z(\cdot\mypp|\CP_n)
=\QQ_z(\cdot)/\QQ_z(\CP_n)$.
\subsection{Statement of the theorem}\label{sec7.1}
As before, we denote $a_z:=\EE_{z}(\xi)$, $K_z:= \Cov(\xi,\xi)$,
$V_z:=K_z^{-1/2}$ \,(see Section \ref{sec5.1}).
Consider the probability density function of a two-dimensional
normal distribution $\mathcal{N}(a_z\myn,K_z)$ \,(with mean $a_z$
and covariance matrix $K_z$), given by
\begin{equation}\label{eq:phi1}
f_{a_z\myn,K_z}(x)= \frac{1}{2\pi\sqrt{\det
K_z\vphantom{^k}}}\:\exp\left(-{\textstyle\frac12}|(x-a_z)\myp
V_z|^2\right),\qquad x\in\RR^2.
\end{equation}
\begin{theorem}\label{th:LCLT}
Assume that $A^+(2)<\infty$ and suppose that Assumption\/
\textup{\ref{as:7.1}} holds. Then\textup{,} uniformly in\/
$m\in\ZZ^2_+$\myp\textup{,}
\begin{equation}\label{eq:LCLT}
\QQ_{z}\{\xi=m\}=f_{a_z\myn,K_z}(m)+O\bigl(|n|^{-5/3}\bigr),\qquad
n\to\infty.
\end{equation}
\end{theorem}
\begin{corollary}\label{cor:Q} Under the conditions of Theorem \textup{\ref{th:LCLT}}
\begin{equation}\label{sim}
\QQ_{z}\{\xi=n\}\asymp (n_1n_2)^{-2/3},\qquad n\to\infty.
\end{equation}
\end{corollary}
Let us point out that the cumulant asymptotics obtained in Section
\ref{sec5.3} (see Lemma \ref{lm:cum}), together with the asymptotics
of the first two moments of $\xi$ (Theorems \ref{th:delta12} and
\ref{th:K}) immediately lead to a central limit theorem
consistent with Theorem \ref{th:LCLT}.
\begin{theorem}[CLT] The distribution of the random vector\/ $(\xi-a_z)\myp
V_z$ converges weakly\textup{,} as $n\to\infty$\textup{,} to the
standard two-dimensional normal distribution $\mathcal{N}(0,I)$.
\end{theorem}
\subsection{Estimates of the characteristic functions}\label{sec7.4}
Before proving Theorem \ref{th:LCLT}, we have to make
some technical preparations.
Recall from Section~\ref{sec2.1} that, with respect to the measure
$\QQ_{z}$, the random variables $\{\nu(x)\}_{x\in \calX}$ are
independent and have characteristic functions (\ref{eq:c.f.}).
Hence, the characteristic function $\varphi_{\xi}(\lambda):=\EE_{z}
(\rme^{\myp\rmi\langle\lambda,\,\xi\rangle})$ of the vector sum
$\xi=\sum_{x\in \calX} x\myp\nu(x)$ is given by
\begin{equation}\label{x.f_5_0}
\varphi_{\xi}(\lambda) =\prod_{x\in \calX} \varphi_{\nu}(\langle
\lambda,x\rangle;x)=\prod_{x\in \calX} \frac{\beta(z^{x}
\rme^{\myp\rmi \langle \lambda,x\rangle})}{\beta(z^{x})}\myp,\qquad
\lambda\in\RR^2.
\end{equation}
Let us start with a general absolute estimate for the characteristic
function of a centered random variable (for a proof, see \cite[Lemma
7.10]{BZ4}).
\begin{lemma}\label{lm:7.2_f}
Let\/ $\varphi_{\nu_0}(t;x):=\EE_{z} (\rme^{\myp\rmi\myp t\myp
\nu_0(x)})$ be the characteristic function of the random variable
$\nu_0(x):=\nu(x)-\EE_{z}[\nu(x)]$. Then
\begin{align}\label{x.f_4}
|\varphi_{\nu_0}(t;x)|\le
\exp\Bigl\{-{\textstyle\frac12}\myp\mu_2(x)t^2+
{\textstyle\frac13}\myp\mu_3(x)|t|^3\Bigr\},\qquad t\in\RR\myp,
\end{align}
where $\mu_q(x):=\EE_{z}|\nu_0(x)^q|$.
\end{lemma}
The next lemma provides two estimates (proved in \cite[Lemmas 7.11
and 7.12]{BZ4})
for the characteristic function
$\varphi_{\xi_0}(\lambda):=\EE_{z}(\rme^{\myp\rmi\langle\lambda,\,\xi_0\rangle})$
of the centered vector $\xi_0:=\xi-a_z
=\sum_{x\in\calX}x\myp\nu_0(x)$. Recall that the Lyapunov
coefficient $L_z$ is defined in \textup{(\ref{L3})}, and
$V_z:=K_z^{-1/2}$.
\begin{lemma}\label{lm:7.2_F}
\textup{(a)} \,For all\/ $\lambda\in\RR^2$\textup{,}
\begin{equation}\label{x.f_6}
|\varphi_{\xi_0}(\lambda V_z)|\le \exp\bigl\{-{\textstyle
\frac12}\myp|\lambda|^2+{\textstyle\frac{1}{3}}\myp
L_z|\lambda|^3\bigr\}.
\end{equation}
\textup{(b)} \,If\/ $|\lambda|\le L_z^{-1}$ then
\begin{equation}\label{16L}
\left|\varphi_{\xi_0}(\lambda V_z)-\rme^{-|\lambda|^2\myn/2}\right|
\le 16\myp L_z|\lambda|^3\mypp \rme^{-|\lambda|^2/6}.
\end{equation}
\end{lemma}
Let us also prove the following global bound (cf.\ \cite[Lemma
7.13]{BZ4}).
\begin{lemma}\label{lm:7.3}
As in Theorem \textup{\ref{th:LCLT}}\textup{,} suppose that
Assumption\/ \textup{\ref{as:7.1}} is satisfied. Then
\begin{equation}\label{f_J}
|\varphi_{\xi_0}(\lambda)|\le\exp\{-C_0 J_{\alpha}(\lambda)\},\qquad
\lambda\in\RR^2,
\end{equation}
where $C_0$ is a positive constant and
\begin{equation}\label{J_0}
J_{\alpha}(\lambda):=\sum_{x\in \calX} \rme^{-\langle\alpha,\myp
x\rangle}\bigl(1-\cos\langle\lambda,x\rangle\bigr)\ge0,\qquad
\lambda\in\RR^2.
\end{equation}
\end{lemma}
\begin{proof}
From (\ref{x.f_5_0}) we have
\begin{equation}\label{eq:J_1}
|\varphi_{\xi_0}(\lambda)|=|\varphi_{\xi}(\lambda)|
=\exp\Biggl\{\sum_{x\in \calX}
\ln\bigl|\varphi_{\nu}(\langle\lambda,\,x\rangle; x)\bigr|\Biggr\}.
\end{equation}
Recall that under Assumption \ref{as:7.1} we have, according to
(\ref{eq:|phi|}) and (\ref{eq:<a1-1}),
\begin{equation*}
\ln|\varphi_{\nu}(t;x)| =\frac12\ln\frac{\beta(z^x\rme^{\myp\rmi\myp
t})\beta(z^x\rme^{-\rmi\myp t})}{\beta(z^x)^2} \le -C_1 b_{1}
z^{x}(1-\cos t),
\end{equation*}
with $C_1>0$ and $b_1>0$. Utilizing this estimate under the sum in
(\ref{eq:J_1}) (with $t=\langle\lambda,\,x\rangle$) and recalling
notation (\ref{alpha}), we arrive at (\ref{f_J}) with $C_0:=C_1
b_1>0$.
\end{proof}
\subsection{Proof of Theorem\/ \textup{\ref{th:LCLT}} and Corollary
\textup{\ref{cor:Q}}}\label{sec7.5}
Let us first deduce the corollary from the theorem.
\begin{proof}[Proof of Corollary \textup{\ref{cor:Q}}] According to
Theorem~\ref{th:4.1}, $a_z=\EE_{z}(\xi)=n+O\bigl(|n|^{2/3}\bigr)$.
Together with Lemma \ref{NV} this implies
\begin{align*}
|(n-a_z)\myp V_z|&\le |n-a_z|\,{\cdot}\, \|V_z\| =O(1).
\end{align*}
Hence, by Lemma \ref{lm:detK} we get
\begin{align*}
f_{a_z\myn,K_z}(n)&=\frac{1}{2\pi\sqrt{\det K_z\vphantom{^k}}}
\:\rme^{-|(n-a_z\myn)V_z|^2\myn/2}\asymp (n_1n_2)^{-2/3},
\end{align*}
and (\ref{sim}) now readily follows from (\ref{eq:LCLT}).
\end{proof}
\begin{proof}[Proof of Theorem \textup{\ref{th:LCLT}}]
By definition, the characteristic function of the random vector
$\xi_0=\xi-a_z$ is given by the Fourier series
\begin{equation*}
\varphi_{\xi_0}(\lambda)
=\sum_{m\in\ZZ_+^2}\QQ_{z}\{\xi=m\}\,\rme^{\myp\rmi\langle
\lambda,\mypp{}m-a_z\myn\rangle},\qquad \lambda\in\RR^2,
\end{equation*}
hence the Fourier coefficients are expressed as
\begin{equation}\label{l_1}
\QQ_{z}\{\xi=m\}=\frac{1}{4\pi^2}\int_{T^2} \rme^{-\rmi\langle
\lambda,\mypp{}m-a_z\myn\rangle}\mypp
\varphi_{\xi_0}(\lambda)\,\dif{}\lambda,\qquad m\in\ZZ_+^2\myp,
\end{equation}
where
$T^2:=\{\lambda=(\lambda_1,\lambda_2)\in\RR^2:|\lambda_1|\le\pi,\,
|\lambda_2|\le\pi\}$. On the other hand, the characteristic function
corresponding to the normal probability density $f_{a_z\myn,K_z}(x)$
(see (\ref{eq:phi1})) is given by
\begin{equation*}
\varphi_{a_z\myn,K_z}(\lambda)=
\rme^{\myp\rmi\langle\lambda,\myp{}a_z\myn\rangle-|\lambda
V_z^{-1}\myn|{\vphantom{(_z}}^2\myn/2},\qquad \lambda\in\RR^2,
\end{equation*}
so by the Fourier inversion formula
\begin{equation}\label{f_o}
f_{a_z\myn,K_z}(m)= \frac{1}{4\pi^2}
\int_{\RR^2}\rme^{-\rmi\langle\lambda,\mypp{}m-a_z\myn\rangle-|\lambda
V_z^{-1}\myn|{\vphantom{(_z}}^2\myn/2}\,\dif{}\lambda\myp,\qquad
m\in\ZZ^2_+\myp.
\end{equation}
Note that if $|\lambda V_z^{-1}\myn|\le L_z^{-1}$ then, according to
Lemmas \ref{NV} and \ref{lm:7.1},
\begin{equation*}
|\lambda|\le |\lambda V_z^{-1}\myn|\cdot\|V_z\| \le L_z^{-1}
\|V_z\|=O\bigl(|n|^{-1/3}\bigr)=o(1),
\end{equation*}
which of course implies that $\lambda\in T^2$. Using this
observation and subtracting (\ref{f_o}) from (\ref{l_1}), we get,
uniformly in $m\in\ZZ^2_+$\myp,
\begin{equation}\label{I}
\bigl|\QQ_{z}\{\xi=m\}-f_{a_z\myn,K_z}(m)\bigr|\le I_1+I_2+I_3\myp,
\end{equation}
where
\begin{align*}
&I_1:=\frac{1}{4\pi^2}\int_{\{\lambda\,:\,|\lambda V_z^{-1}\myn|
\le L_z^{-1}\}}
\bigl|\varphi_{\xi_0}(\lambda)-\rme^{-|\lambda V_z^{-1}\myn|{\vphantom{(_z}}^2\myn/2}
\bigr|\,\dif{}\lambda\myp,\\
&I_2:=\frac{1}{4\pi^2}\int_{\{\lambda\,:\,|\lambda V_z^{-1}\myn|>
L_z^{-1}\}}
\rme^{-|\lambda V_z^{-1}\myn|{\vphantom{(_z}}^2\myn/2}\,\dif{}\lambda\myp,\\
&I_3:=\frac{1}{4\pi^2}
\int_{T^2\cap\{\lambda\,:\,|\lambda V_z^{-1}\myn|>L_z^{-1}\}}
|\varphi_{\xi_0}(\lambda)|\:\dif{}\lambda\myp.
\end{align*}
By the substitution $\lambda=y\myp V_z$, the integral $I_1$ is
reduced to
\begin{equation}\label{I1}
\begin{aligned}
I_1&=\frac{|\mynn\det V_z|}{4\pi^2} \int_{|y|\le L_z^{-1}}
\bigl|\varphi_{\xi_0}(y V_z)-
\rme^{-|y|^2\myn/2}\bigr|\,\dif{}y\\
&=O(1)\mypp(\det K_z)^{-1/2}\myp L_z\int_{\RR^2}
|y|^3\rme^{-|y|^2\myn/6}\,\dif{}y= O(|n|^{-5/3}),
\end{aligned}
\end{equation}
on account of Lemmas \ref{lm:detK}, \ref{lm:7.1} and
\ref{lm:7.2_F}(b). Similarly, again putting $\lambda=y\myp V_z$ and
passing to the polar coordinates, we get, due to Lemmas
\ref{lm:detK} and \ref{lm:7.1},
\begin{equation}\label{I2}
\begin{aligned}
I_2&= \frac{|\mynn\det V_z|}{2\pi} \int_{L_z^{-1}}^\infty r\mypp
\rme^{-r^2\myn/2}\,\dif{}r=O(|n|^{-4/3})\,
\rme^{-L_z^{-2}\myn/2}=o(|n|^{-5/3}).
\end{aligned}
\end{equation}
Finally, let us turn to $I_3$. Using Lemma \ref{lm:7.3}, we obtain
\begin{align}\label{I3}
I_3&= O(1)\int_{T^2\cap \{|\lambda V_z^{-1}\myn|>L_z^{-1}\}}
\rme^{-C_0J_{\alpha}(\lambda)}\,\dif{}\lambda\myp,
\end{align}
where $J_{\alpha}(\lambda)$ is given by (\ref{J_0}).
$|\lambda V_z^{-1}\myn|>L_z^{-1}$
then $|\lambda|>\eta\myp|\alpha|$ for a suitable (small enough)
constant $\eta>0$, which implies that
$\max\{|\lambda_1|/\alpha_1,|\lambda_2|/\alpha_2\}>\eta$, for
otherwise from (\ref{alpha}) and Lemmas \ref{lm:K_z} and
\ref{lm:7.1} it would follow
\begin{align*}
1<L_z|\lambda V_z^{-1}\myn|&\le L_z\myp\eta\myp|\alpha|\,{\cdot}\,
\|K_z\|^{1/2}= O(\eta)\to0\quad\text{as}\ \ \eta\downarrow0,
\end{align*}
which is a contradiction. Hence, estimate (\ref{I3}) is reduced to
\begin{align}\label{I3.1}
I_3&=O(1) \left(
\int_{|\lambda_1|>\eta\myp\alpha_1}+\int_{|\lambda_2|>\eta\myp\alpha_2}\right)
\rme^{-C_0J_{\alpha}(\lambda)}\,\dif{}\lambda.
\end{align}
To estimate the first integral in (\ref{I3.1}), by keeping in
summation (\ref{J_0}) only pairs of the form $x=(x_1,1)$,
\,$x_1\in\ZZ_+$\myp, we obtain
\begin{align}
\notag J_{\alpha}(\lambda)
\ge \sum_{x_1=0}^\infty \rme^{-\alpha_1
x_1}\!\left(1-\Re\,
\rme^{\myp\rmi (\lambda_1 x_1+\lambda_2)}\right)
&=\frac{1}{1-\rme^{-\alpha_1}}-
\Re\left(\frac{\rme^{\myp\rmi\lambda_2}}{1-\rme^{- \alpha_1+\rmi\lambda_1}}\right)\\
\label{J_1}& \ge
\frac{1}{1-\rme^{-\alpha_1}}-\frac{1}{|1-\rme^{-\alpha_1 +
\rmi\lambda_1}|}\myp,
\end{align}
because $\Re\myp u\le |u|$ for any $u\in\CC$. Since
$\eta\mypp\alpha_1\le|\lambda_1|\le \pi$, we have
\begin{align*}
|1-\rme^{-\alpha_1+\rmi\lambda_1}|
&\ge|1-\rme^{-\alpha_1+\rmi\myp\eta\myp\alpha_1}|\sim\alpha_1
(1+\eta^2)^{1/2} \qquad (\alpha_1\to0).
\end{align*}
Substituting this estimate into (\ref{J_1}), we conclude that
$J_\alpha(\lambda)$ is asymptotically bounded from below by
$C(\eta)\mypp\alpha_1^{-1}\mynn \asymp |n|^{1/3}$ (with some
constant $C(\eta)>0$), uniformly in $\lambda$ such that
$\eta\mypp\alpha_1\le|\lambda_1|\le \pi$. Thus, the first integral
in (\ref{I3.1}) is bounded by
\begin{equation*}
O(1)\exp\bigl(-\myp\const \cdot
|n|^{1/3}\bigr)=o\bigl(|n|^{-5/3}\bigr).
\end{equation*}
Similarly, the second integral in (\ref{I3.1}) is estimated by
reducing the summation in (\ref{J_0}) to that over $x=(1,x_2)$ only.
As a result, $I_3=o\bigl(|n|^{-5/3}\bigr)$. Substituting this
estimate, together with (\ref{I1}) and (\ref{I2}), into (\ref{I}) we
get (\ref{eq:LCLT}), and so the theorem is proved.
\end{proof}
\section{Proof of the limit shape results}\label{sec8}
Let us first establish the universality of the limit shape under the
measure
$\QQ_z$.
\begin{theorem}\label{th:8.2}
For each\/ $\varepsilon>0$\textup{,}
\begin{equation*}
\lim_{n\to\infty} \QQ_{z}\biggl\{\sup_{0\le t\le\infty}\bigl|\myp
n_j^{-1}\xi_j(t)-g^*_j(t)\bigr| \le\varepsilon\biggr\}=1\qquad
(j=1,2).
\end{equation*}
\end{theorem}
\begin{proof}
By Theorems \ref{th:3.2} and \ref{th:8.1.1a}, the expectation of the
random process $n_j^{-1}\xi_j(t)$ uniformly converges to $g^*_j(t)$
as $n\to\infty$. Therefore, we only need to check that, for each
$\varepsilon>0$,
\begin{equation*}
\lim_{n\to\infty}\QQ_{z}\biggl\{\sup_{0\le t\le\infty}
n_j^{-1}\bigl|\myp\xi_j(t)-\EE_{z}[\xi_j(t)]\bigr|>\varepsilon\biggr\}=0.
\end{equation*}
Note that the random process
$\xi_{0j}(t):=\xi_j(t)-\EE_{z}[\xi_j(t)]$
is a martingale with respect to the
filtration ${\mathcal F}_t:=\sigma\{\nu(x),\, x\in \calX(t)\}$,
\,$t\in[0,\infty]$. From the definition of $\xi_j(t)$ (see
(\ref{xi(t)})), it is also clear that $\xi_{0j}(t)$ is
a c\`adl\`ag process (i.e., its paths are everywhere
right-continuous and have left limits).
Therefore, applying the Kolmogorov--Doob submartingale inequality
(see, e.g., \cite[Corollary~2.1, p.\,14]{Yeh}) and using Theorem
\ref{th:K}, we obtain
\begin{align*}
&\QQ_{z}\biggl\{\sup_{0\le t\le\infty}|\myp\xi_{0j}(t)|> \varepsilon
n_j\biggr\} \le\frac{\Var(\xi_j)}{(\varepsilon n_j)^2}=
O\bigl(|n|^{-2/3}\bigr)\to0,
\end{align*}
and the theorem is proved.
\end{proof}
We are finally ready to prove our main result about the universality
of the limit shape under the measures $\PP_n$ (cf.\
Theorem~\ref{th:main1}).
\begin{theorem}\label{th:8.2a}
For any\/ $\varepsilon>0$\textup{,}
\begin{equation*}
\lim_{n\to\infty} \PP_n\biggl\{\sup_{0\le t\le\infty}\bigl|\myp
n_j^{-1}\xi_j(t)-g^*_j(t)\bigr| \le\varepsilon\biggr\}=1\qquad
(j=1,2).
\end{equation*}
\end{theorem}
\begin{proof}
Like in the proof of Theorem \ref{th:8.2}, the claim is reduced to
the limit
\begin{equation}\label{A1-0}
\lim_{n\to\infty} \PP_n\biggl\{\sup_{0\le t\le\infty}
\bigl|\myp\xi_{0j}(t)\bigr|>\varepsilon n_j\biggr\}=0,
\end{equation}
where $\xi_{0j}(t):=\xi_j(t)-\EE_{z}[\xi_j(t)]$. Using (\ref{Pn}) we
get
\begin{align}\label{A1}
\PP_n\biggl\{\sup_{0\le t\le\infty}\bigl|\myp\xi_{0j}(t)\bigr|>
\varepsilon n_j\biggr\} \le \frac{\QQ_{z}\!\left\{\sup_{0\le
t\le\infty}\bigl|\myp\xi_{0j}(t)\bigr|> \varepsilon
n_j\right\}}{\QQ_{z}\{\xi=n\}}\myp.
\end{align}
By the Kolmogorov--Doob submartingale inequality and Lemma
\ref{lm:6.7} (with $q=4$), we have
\begin{equation}\label{A1-1}
\QQ_{z}\biggl\{\sup_{0\le
t\le\infty}\bigl|\myp\xi_{0j}(t)\bigr|>\varepsilon n_j \biggr\}\le
\frac{\EE_{z}\bigl|\myp\xi_j-\EE_z(\xi_j)\bigr|^{4}}{(\varepsilon
n_j)^{4}} =O\bigl(|n|^{-2}\bigr).
\end{equation}
On the other hand, by Corollary \ref{cor:Q}
\begin{equation}\label{A1-2}
\QQ_{z}\{\xi=n\}\asymp(n_1n_2)^{-2/3}\asymp |n|^{-4/3}.
\end{equation}
Combining (\ref{A1-1}) and (\ref{A1-2}), we conclude that the
right-hand side of (\ref{A1}) is dominated by a quantity of order of
$O\bigl(|n|^{-2/3}\bigr)\to0$, and so the limit in (\ref{A1-0})
follows.
\end{proof}
\section*{Acknowledgements}
This work was supported in part by a Leverhulme Research Fellowship.
The author is grateful to \mypp{}B.~L.~Granovsky, \,S.\,A.~Molchanov
and Yu.\,Yakubovich for helpful discussions.
|
\section{Introduction}
In recent years there has been a growing interest in the study of
strongly correlated quantum states and their possible realization
in two-dimensional (2D) systems of ultracold
atoms~\cite{Lewenstein07,Bloch08}. Such states, which were first
postulated when studying the dynamics of electrons subjected to
strong magnetic fields, can also be produced in systems of neutral
atoms subjected to so-called artificial gauge fields. One of the
first examples was obtained by rotating a 2D atomic cloud such that
the centrifugal force on the atoms mimics the Lorentz force which a
charged particle would experience in the presence of a constant
magnetic field perpendicular to the
system~\cite{Wilkin:2000,fetter,cooper,Dagnino:2009}. The main
drawback of this approach is that large rotations are needed in order
to observe strongly correlated states such as the
Laughlin~\cite{Laughlin:1983}, while it is difficult to stabilize
in this fast rotating regime. For this reason, the Laughlin state
has not yet been engineered in the pioneering
experiments ~\cite{Bretin:2004,schweikhard}. Recently, Roncaglia
{\textit et al.} have proposed an alternative experiment to avoid the
instability difficulty by using a Mexican-hat trap~\cite{roncaglia}.
Other very promising proposals to overcome this problem come from
quantum optics and consider the coupling of the atoms to one or
several laser fields. These make the atoms experience a Berry
phase~\cite{oehberg,Dalibard:2011,ours11,ours2,levi}, which, due to
the mathematical equivalence between geometric phases and
external gauge fields, can then be interpreted as if it were
due to the presence of an external gauge field. The experimental
realization of such an artificial gauge field has already been
achieved~\cite{lin}.
An important motivation to study these systems is the
possibility of producing strongly correlated quantum states and
quasi-particle/quasi-hole excitations which are neither
described by fermionic nor by bosonic commutation laws. The
latter are expected to have strong impact in the context of
anyonic quantum computation~\cite{nayak}. One thus needs to
quantify the strongly correlated states produced by different
proposals and the properties of their quasi-particle/quasi-hole
excitations.
A common way to study such systems theoretically is by means
of exact diagonalization of small-sized systems~\cite{Wilkin:2000,Dagnino:2009}.
The usual methodology is to employ Fock-Darwin (FD) wave functions,
which describe single particles with fixed angular momentum in
a fixed Landau level, and which are the eigenfunctions of a 2D
non-interacting system with a perpendicular magnetic field in
the symmetric gauge. The many-body basis for bosons (fermions)
is then built up by symmetric (antisymmetric) combinations of
the FD states. While this basis is practical for definite
calculations, many relevant states in the literature have been
found by proposing a first-quantized, analytic wave function.
Here the Laughlin~\cite{Laughlin:1983}, the Pfaffian, also
called Moore-Read~\cite{Moore:1991}, or the Laughlin
quasi-particle states~\cite{Popp:2004} are the most prominent
examples. Translating the first-quantized wave functions into
the language of second quantization, however, turns out to be
a hard task~\cite{takano,mitra}.
In this paper, we present a computer code which achieves
this goal for arbitrary states described by an analytic function,
and thus provides practitioners of this field with a simple
and yet powerful tool to quantify the degree of correlation
by examining its expansion into an independent particle motion
basis. The code is written in Mathematica~\cite{Mathematica},
which is a computer language specially suited for symbolic
evaluation.
We begin by presenting the first-quantized expression of the
most important strongly correlated states in Sect.~\ref{sec:st}.
In Sect.~\ref{sec1}, we briefly describe and construct the
many-body basis into which we then decompose the states in
Sect.~\ref{sec:deco}. Finally, in Sect.~\ref{sec:app}, we
consider two applications which can be tackled making use of
the described decomposition scheme. The most relevant routines
contained in {\bf Strongdeco.nb} are explained within the
text, a brief description of all routines is given in the Appendix.
\section{Analytical strongly correlated states}
\label{sec:st}
Strongly correlated states in 2D systems exposed to a gauge field
are usually studied in the regime where all particles occupy
the lowest Landau level. The Hilbert space of an $N$-body system
in this regime can be represented by wave functions of the form
\begin{equation}
\Psi(z_1,\dots,z_N) = {\mathcal{N}} f(z_1,\dots,z_N) {\mathrm e}^{-\sum \mid
z_i\mid^2/ 2
\lambda_{\perp}^2}\, ,
\label{form}
\end{equation}
where $z_i=(x_i+i y_i)/\lambda_{\perp}$, ${\mathcal N}$ is a normalization
constant, and $f$ is a polynomial in its arguments $z_i$. The typical
length scale of the system is given by $\lambda_{\perp}$. The most
famous wave function of this form is the Laughlin
function~\cite{Laughlin:1983}:
\begin{equation}
\Psi_{\mathcal L}(z_1,\dots,z_N) ={\mathcal{N}_{\mathcal L}} f_m(z_i,\dots,z_N)
{\mathrm e}^{-\sum \mid z_i\mid^2/ 2
\lambda_{\perp}^2}\, ,
\label{laughwave}
\end{equation}
with ${\mathcal N}_{\mathcal L}$ a normalization constant, and
\begin{equation}
f_m(z_i,\dots,z_N) \equiv \prod_{i<j}(z_i-z_j)^m\,,
\end{equation}
where $m$ is an integer directly related to the filling
factor $\nu=1/m$ of the lowest Landau level. Originally
intended to describe electrons, this wave function had
to be antisymmetric, restricting $m$ to odd numbers. However,
as shown for instance in Ref.~\cite{Wilkin:2000}, also the
ground state of a two-dimensional system of rotating bosons
with contact interaction is, for certain values of the angular
rotation, described by the Laughlin state, if $m$ is taken as an
even integer~\cite{Cooper:2001,Regnault:2003}. One important
property of the Laughlin wave function is that $f_m$ is a
homogeneous polynomial. Its degree determines the well-defined
total angular momentum of the system, given by $L=\frac{1}{2}mN(N-1)$.
Besides the Laughlin state, other states of the form given
by Eq.~(\ref{form}) show up as the ground state of a rotating
Bose gas, if we vary the rotation frequency~\cite{Wilkin:2000,Popp:2004,Roncaglia:2010}.
For a broad range of rotation frequencies, for instance, a large
overlap is found with the so-called Pfaffian state, which has
$L=N(N-2)/2$ for even $N$, and $L=(N-1)^2/2$ for odd $N$. It
explicitly reads,
\begin{equation}
\Psi_{\mathcal P}= {\mathcal N}_p {\rm Pf}([z]) \prod_{i<j}(z_i-z_j) \ ,
\label{pfaf}
\end{equation}
with ${\mathcal N}_p$ a normalization coefficient and,
\begin{equation}
{\rm Pf}([z])= {\mathcal A} \left[ {1 \over (z_1-z_2)} {1 \over (z_3-z_4)} \cdots {1
\over (z_{N-1}-z_N)}\right]{\mathrm e}^{-\sum \mid z_i\mid^2/ 2
\lambda_{\perp}^2}\, ,
\end{equation}
where ${\mathcal A}$ is an antisymmetrizer of the product. As explained
in Ref.~\cite{Wilkin:2000}, the Pfaffian state can also be computed
as,
\begin{equation}
\Psi_{\mathcal P}=
{\mathcal S}
\prod_{i<j \in \sigma_1} (z_i-z_j)^2
\prod_{k<l \in \sigma_2}(z_k-z_l)^2 \;
{\mathrm e}^{-\sum \mid z_i\mid^2/ 2\lambda_{\perp}^2}\, ,
\label{conje}
\end{equation}
where $\sigma_1$ and $\sigma_2$ are two subsets containing $N/2$
particles each if $N$ is even, and $(N+1)/2$ and $(N-1)/2$ if
$N$ is odd. ${\mathcal S}$ symmetrizes the expression.
Another relevant state is the Laughlin quasi-particle state,
\begin{equation}
\Psi_{{\mathcal L}qp}(\xi,\xi^*)=
{\mathcal N}_{qp}(\xi,\xi^*) {\mathrm e}^{-\sum \mid z_i\mid^2/ 2
\lambda_{\perp}^2} \prod_{i\leq N} (\partial_{z_i}-\xi) f_m\,,
\label{lqp}
\end{equation}
where $\xi$ represents the position of the quasiparticle. If
we pin the quasi-particle to the origin, the Laughlin
quasi-particle state has a definite angular momentum
$L=\frac{1}{2}mN(N-1)-N$, and, as shown in Ref.~\cite{Popp:2004}
for $m=2$ and $N=4$, its overlap with the ground state of
rotating ultracold atoms is fairly large in certain regions
of the rotation.
The analog of Laughlin's quasi-particle state is the quasi-hole
state,
\begin{equation}
\Psi_{{\mathcal L}qh}(\xi,\xi^*)=
{\mathcal N}_{qh}(\xi,\xi^*){\mathrm e}^{-\sum \mid z_i\mid^2/ 2
\lambda_{\perp}^2} \prod_{i\leq N} (z_i-\xi) f_m\,,
\label{lqh}
\end{equation}
with an increased angular momentum
$\frac{1}{2}mN(N-1) \leq L\leq \frac{1}{2}mN(N-1)+N$. An interesting
property of quasiparticles and quasiholes is their anyonic
nature~\cite{arovas} and fractional charge. Note that in the
case of an electroneutral system, one may define the analog
of a charge by looking at the Berry phase a particle acquires
when moving in the presence of the artificial gauge field.
\section{The many-body basis}
\label{sec1}
After defining in the previous section the general structure
for all wave functions of interest, we now construct the
many-body basis into which we wish to decompose these states.
A convenient choice to span the Hilbert space of the many-body
system are the eigenfunctions of the non-interacting problem,
i.e. the FD wave functions,
\begin{equation}
\phi_{\ell} (z) =
{z^\ell \over \sqrt{\pi \ell!}}
{1\over \lambda_\perp^{\ell + 1}}
\;
e^{-|z|^2/(2\lambda_\perp^2)} \,, \quad \ell=0,\dots,\infty \, ,
\label{fd}
\end{equation}
where we have restricted ourselves to the lowest Landau level.
These states satisfy,
\begin{equation}
\int_{-\infty}^\infty \,dx
\int_{-\infty}^\infty \,dy\,
\phi^*_{\ell}(z) \phi_{\ell'}(z) =\delta_{\ell,\ell'} \,.
\end{equation}
The many-body basis can be generated by considering products of
the FD functions, which in the case of bosons have to be combined
in a symmetric way, while antisymmetric combinations must be
constructed in the case of fermions. Here, we will concentrate
on the bosonic case, but with only slight modifications which
are explicitly shown in the code file, fermionic systems can be
treated in the same way. For the bosonic system we write the
many-body state as,
\begin{equation}
\{\ell_1, \ell_2, \dots, \ell_N\} \equiv
{\mathcal S} \;\left[
\phi_{\ell_1}(z_1)
\phi_{\ell_2}(z_2)
\dots
\phi_{\ell_N}(z_N)\right]
\end{equation}
where ${\mathcal S}$ symmetrizes over the $N$ particles. These states
are called permanents, which are the bosonic analog of the Slater
determinants, with the difference that all terms have a positive
sign. Without loss of generality we may assume that
$\ell_1 \leq \ell_2 \leq \cdots \leq \ell_N$. The orthonormality
of the permanents then reads
\begin{equation}
\label{onb}
\{\ell_1, \ell_2, \dots, \ell_N\}\cdot\{\ell_1', \ell_2', \dots, \ell_N'\} =
\delta_{\ell_1,\ell_1'}\delta_{\ell_2,\ell_2'} \cdots \delta_{\ell_N,\ell_N'}\,.
\end{equation}
For simplicity we will from now on set the scale factor $\lambda_\perp=1$,
and suppress the exponential term which is common to all $N$-body
states, and, as an overall Gaussian, fixes the center of mass to the
origin. We can then simplify the problem to dealing with permanents
of the form,
\begin{equation}
{\mathcal S}\left[z_1^{\ell_1} z_2^{\ell_2}\dots z_N^{\ell_N} \right]\,.
\end{equation}
From Eq.~(\ref{onb}) follows that, for a given $N$, all states of a
fixed total angular momentum $L=\sum_{i=1}^N \ell_i$ form a subspace
which is orthogonal to the subspace with total angular momentum
$L' \neq L$. We can therefore perform the decomposition independently
in each subspace, and thus restrict ourselves to a subspace with
fixed $L$. Its basis (up to normalization factors and the overall
exponential term) can be constructed through the command,
\vspace{10pt}
\begin{boxedminipage}{0.8\textwidth}
\begin{verbatim}
ConjS[na_, L_] := Module[{poty, dimy},
poty = Pots[na, L];
dimy = Dimensions[poty][[1]];
Table[Perm[na, poty[[i]]], {i, 1, dimy}]]
\end{verbatim}
\end{boxedminipage}
\vspace{10pt}
\noindent which makes use of the function {\bf Perm}~\cite{Eric},
that builds the appropriate permanent, and of {\bf Pots[N,L]}, which
constructs the set of indexes ${\ell_1, \dots \ell_N}$ for a given
$N$ and $L$, represented by {\bf na} and {\bf L} in the code,
\vspace{10pt}
\begin{boxedminipage}{0.94\textwidth}
\begin{verbatim}
cc[0] = 0;
tab[n_, l_] :=
Table[{cc[i], cc[i - 1],
If[i == 1, l, (l - Sum[cc[j], {j, 0, i - 1}])/2]},
{i, 1, n - 1}];
Pots[na_, L_] := If[na == 2, Table[{i, L - i}, {i, 0, L/2}],
Module[{pat},
Clear[pat];
pat[na] = Join[Table[
cc[i], {i, 1, na - 1}], {L - Sum[cc[i], {i, 1, na - 1}]}];
pat[a_] := Table[pat[a + 1], Evaluate[tab[na, L][[a]]]];
Flatten[pat[1], na - 2]]]
\end{verbatim}
\end{boxedminipage}
\vspace{10pt}
\noindent For instance, for $N=4$ and $L=2$ we have
~\footnote{Note that the state $\{1,1,0,0\}$ is equivalent to $\{0,0,1,1\}$
due to the symmetrization of the states.},
\begin{verbatim}
Pots[4,2]={{0,0,0,2},{0,0,1,1}}
\end{verbatim}
and correspondingly,
\begin{verbatim}
ConjS[4, 2]=
{6 z[1]^2+6 z[2]^2+6 z[3]^2+6 z[4]^2,
4 z[1] z[2]+4 z[1] z[3]+4 z[2] z[3]
+4 z[1] z[4]+4 z[2] z[4]+4 z[3] z[4]}
\end{verbatim}
As can be seen in this example, due to multiple occupation
of the same single-particle state, some of the permutations
contributing to the symmetrized wavefunction are described
by the same monomials which thus have prefactors given
by the factorial of the number of permutations. These
factors need to be taken into account to correctly normalize
the many-body states, and can be obtained through, {\bf nami[N, L]},
which gives a table with the same ordering as {\bf Pots} or
{\bf ConjS}, for our previous example, {\bf nami[4,2]=\{6,4\}},
as could be inferred from the obtained expressions.
\vspace{10pt}
\begin{boxedminipage}{0.94\textwidth}
\begin{verbatim}
nami[na_, L_] := Module[{potty, pp, inde, ta},
potty = Pots[na, L];
pp = Dimensions[potty][[1]];
inde = Table[Complement[potty[[i]]], {i, 1, pp}];
ta = Table[Table[Count[potty[[i]], inde[[i, j]]],
{j, 1, Dimensions[inde[[i]]][[1]]}], {i, 1, pp}];
Table[ Product[ta[[i, j]]!,
{j, 1, Dimensions[ta[[i]]][[1]]}] , {i, 1, pp}]]
\end{verbatim}
\end{boxedminipage}
\vspace{10pt}
\noindent Once these factors are known it is easy to build
the normalization coefficient by looking into the prefactors
in the Fock-Darwin states Eq.~(\ref{fd}), $1/\sqrt{\pi \ell!}$.
The function {\bf tip[N, L]} gives the normalization
coefficients. Their explicit coding is,
\vspace{10pt}
\begin{boxedminipage}{0.95\textwidth}
\begin{verbatim}
tip[na_, L_] := Module[{potty, nimy},
potty = Pots[na, L];
nimy = nami[na, L];
Table[Sqrt[nimy[[i]]]Sqrt[Product[Pi Gamma[potty[[i, jj]]+1],
{jj, 1, na}] ], {i, 1, Dimensions[nimy][[1]]}]]
\end{verbatim}
\end{boxedminipage}
\section{Decomposition of the states}
\label{sec:deco}
All states described in Sect.~\ref{sec:st}, are, up to
the common exponential factor, polynomials in the $z$
variables. To write down the states in terms of the
many-body ones, we can suppress the exponential and work
out the decomposition of the polynomial in terms of the
permanents. While the Laughlin and the Pfaffian state have
a definite total angular momentum, for the quasi-hole and
quasi-particle states this is only true if we fix the
position of the quasi-particle to the origin $\xi=0$. Otherwise,
we must first sort the polynomial by the different contributions
with a definite order in $z$, and can then proceed, for each
contribution separately, in the way described here, where
we assume an analytical state, $\Psi(z_1,z_2,\dots,z_N)$
with fixed $N$ and $L$. We are looking for an expansion of
the form,
\begin{equation}
\label{dec}
\Psi(z_1,z_2,\dots,z_N) = \sum_{j=1}^{n_D} C_j \{
\ell_{1,j},\ell_{2,j},\dots, \ell_{N,j}\}
\end{equation}
where $n_D$ is the total size of the many-body basis,
which can be computed as $n_D={\rm Dimensions[PotsN}[N,L]][[1]]$.
To get a feeling of how this grows with $N$ and $L$ the dimension
of these spaces for the $L$ corresponding to the Laughlin wave
functions are, $n_D=7, 34, 192, 1206, 8033, 55974$ for
$N=3,4,5,6,7$ and 8, respectively.
To decompose a polynomial into these states, we have to find
the monomials which correspond to a given Fock state and read
out their coefficients. Since we know that the polynomial is
symmetric (antisymmetric) under exchange of two coordinates,
it is sufficient to find only one monomial contributing to a given
Fock state, as all the others must have the same coefficient (up to a sign in
the antisymmetric case). This can be achieved by taking derivatives:
\begin{equation}
\partial_{z_1}^{\ell_{1,j}} \cdots \partial_{z_N}^{\ell_{N,j}}
\Psi(z_1,z_2,\dots,z_N) |_{z_1=0,\cdots,z_N=0}
= c_j,
\label{eq15}
\end{equation}
where $c_j$ is not yet the coefficient $C_j$ in Eq. (\ref{dec}), but is directly
related to it through the normalization procedure described in the previous
section. Hereby, we have to take into account that an additional factor
$\prod_{i=1}^{N} \ell_{i,j}!$ occurs through the derivatives. Thus we
obtain
\begin{equation}
C_j = c_j \left(P \prod_{i=1}^{N} \ell_{i,j}!
\ \pi \right)^{-1/2} \equiv
\alpha_j c_j,
\end{equation}
where the $P$ is the factorial of permutations leading to the same expression,
obtained by {\bf nami[N,L]}. We thus see that $\alpha_j$ equals the inverse
of the $j$th component of ${\bf tip[N,L]}$. The decomposition of, for instance,
the Laughlin wavefunction can therefore be obtained by the following piece of
code:
\vspace{10pt}
\begin{boxedminipage}{0.97\textwidth}
\begin{verbatim}
DDecoLaug[na_,nu_] :=
Module[{Lmin, Lmax, state, base, dim, factors, d, prf, outp},
Lmin = na (na - 1);
Lmax = na (na - 1);
state = Laughlin[na,nu];
base = Flatten[Table[Pots[na, i], {i, Lmin, Lmax}], 1];
dim = Dimensions[base][[1]];
factors = Flatten[Table[tip[na, i], {i, Lmin, Lmax}], 1];
d[0] = state;
prf = Table[
For[i = 1, i < na + 1, i++,
d[na] = 0;
d[i] = D[d[i - 1], {z[i], base[[j, i]]}];
d[i] = d[i] /. z[i] -> 0;
If[d[i] == 0, Break[]]
];
d[na]/factors[[j]],
{j, 1, dim}];
outp = prf/Sqrt[prf.prf]]
\end{verbatim}
\end{boxedminipage}
\vspace{10pt}
\noindent Here, {\bf Laughlin[N,nu]} describes the Laughlin wavefunction
for $N$ particles at filling $\nu$. For even $1/\nu$, this is a symmetric
function describing bosons, while odd values yield an antisymmetric
function for fermionic systems. In principle, we can use the code
for both the symmetric and the antisymmetric case. In the latter,
however, it is convenient to exclude states with multiple occupied
single-particle levels from the basis, as they obviously make no
contribution. This can be done by replacing {\bf Pots[N,L]} by its
fermionic analogue {\bf PotsF[N,L]} defined in the code file.
Consequently, we will also have to replace {\bf tip[N,L]} by
{\bf tipF[N,L]}.
An alternative way to achieve the decomposition is by
means of a particular, built-in Mathematica function,
{\bf PolynomialReduce}. This function provides the decomposition
of a given multivariate polynomial in terms of a set of
polynomials. The code for decomposing the bosonic Laughlin
state then reads
\vspace{10pt}
\begin{boxedminipage}{0.97\textwidth}
\begin{verbatim}
LaugDeco[na_,nu_]:=Module[{state, base, symb, laur, prf, outp},
state = Laughlin[na,nu];
base = ConjS[na, na (na - 1)];
symb = Table[z[i], {i, 1, na}];
laur = PolynomialReduce[state, base, symb];
If[laur[[2]] != 0, Print["Problem in reduction"]];
prf = laur[[1]] tip[na, na (na - 1)];
outp = prf/Sqrt[prf.prf];
outp]
\end{verbatim}
\end{boxedminipage}
\vspace{10pt}
\noindent For most states that we have considered, the
decomposition by means of derivatives is faster. However,
making use of {\bf PolynomialReduce} turns out to be
quicker for the fermionic Laughlin state as well as
for quasiparticle excitations.
In figure~\ref{figy}, a snapshot of the code
for the decomposition of the Laughlin state is provided
for $N=3$, $N=4$ and $N=5$. The code has been tested for $N\leq 7$
on a laptop running on linux with 1Gb of RAM memory. A listing
of the different commands defined in {\bf Strongdeco.nb} is
provided in the Appendix. The notebook is provided with some
examples built-in inside.
\begin{figure}[h]
\fbox{\parbox{1.0\textwidth}{\includegraphics[width=12.cm]{fig1.eps}}}
\caption{Snapshot of the code where the decomposition of the
Laughlin state with $\nu=1/2$ is obtained for $N=3$, $N=4$ and $N=5$. }
\protect\label{figy}
\end{figure}
\section{Applications}
\label{sec:app}
Finally, we make use of the presented decomposition scheme,
and consider as examples three problems which might be
tackled with the given code.
\subsection{Wave-function overlaps}
The decomposition achieved by our code turns out to be very
useful if the state, e.g. the eigenstates of a certain problem
or the evolved state at a given time, of a system is known
in the many-body basis. This is the case if a system is
studied via exact diagonalization ~\cite{Wilkin:2000,Dagnino:2009,ours11,ours2}.
It is then customary to ask whether, and to which degree, the
obtained state resembles one of the well-known strongly
correlated states described in Sect.~\ref{sec:st}. To answer
this question one has to calculate the overlap between both
states, which can either be done by straightforwardly expressing
the many-body state in first quantization and then evaluating the
overlap integrals. For reasonable system sizes of $N \geq 3$,
the latter is a very lengthy task. The second possibility
consists of finding an expression of the analytic wave function
in terms of the many-body basis, which then reduces the overlap
calculation to a simple scalar multiplication of two vectors.
In this case, the first step is the non-trivial one, but it is
directly achieved by the code we have presented here.
\subsection{Angular-momentum distribution}
As a second application, one can consider a system which is known
to have eigenstates of the form given by Eq.~(\ref{form}). Many of
the properties of such states are better computed by first transforming
it into the independent motion basis. A clear example is the calculation
of the angular momentum distribution of the state, from which one also
gains insight into its one-body density matrix and other correlation
functions. For the fermionic Laughlin state this problem has been
considered in Ref.~\cite{mitra}, where exact results are obtained by
calculating the density and then extracting the angular momentum
distribution. This method, however, fails for systems larger than
$N=3$, for which MonteCarlo methods have been applied. By means of
our code, we are able to reproduce these results by decomposing the
Laughlin state into a basis from which the angular momentum of each
particle can be directly read off. It is straightforward to go beyond the
analytical results of Ref. \cite{mitra}. As an example,
we give in Table~\ref{ad} the angular momentum distributions for
the fermionic Laughlin state at $m=3$ and the bosonic Laughlin
state at $m=2$ of a system with $N=4$ particles.
\begin{table}
\begin{center}
\begin{tabular}{|l|cccccccccc|}
\hline
$\ell=$ & $0$ & $1$ & $2$ & $3$ & $4$ & $5$ &
$6$ & $7$ & $8$ & $9$ \\
\hline
& \multicolumn{10}{l|}{}\\
$m=2$ & $\frac{257}{553}$ & $\frac{264}{553}$& $\frac{303}{553}$&
$\frac{446}{553}$& $\frac{447}{553}$& $\frac{330}{553}$& $\frac{165}{553}$&
0& 0& 0\\
& \multicolumn{10}{l|}{}\\
$m=3$ & $\frac{185}{706}$ & $\frac{185}{706}$& $\frac{209}{706}$&
$\frac{321}{706}$& $\frac{417}{706}$& $\frac{465}{706}$& $\frac{455}{706}$&
$\frac{339}{706}$& $\frac{186}{706}$& $\frac{62}{706}$\\
& \multicolumn{10}{l|}{}\\
\hline
\end{tabular}
\caption{\label{ad} Angular-momentum distribution for Laughlin states of $N=4$
particles for bosons ($m=2$) and fermions ($m=3$).}
\end{center}
\end{table}
\subsection{Fractional charge of excitations}
Another useful application is the calculation of the
normalization factor for a state of the form (\ref{form}).
As explained in Ref.~\cite{wen}, the normalization factors
${\mathcal N}_{qh}(\xi,\xi^*)$ and ${\mathcal N}_{qp}(\xi,\xi^*)$ of the
quasi-particle in Eq.~(\ref{lqp}) and the quasi-hole state in
Eq.~(\ref{lqh}), contain information about the Berry phase
${\bf a}\cdot\mathrm{d}{\bf x}$ which these excitations acquire
during an adiabatic movement:
\begin{equation}
a_{\xi} \equiv \frac{i}{2} \partial_{\xi} \mathrm{ln} {\mathcal N}(\xi,\xi^*),
\end{equation}
with $\partial_{\xi}\equiv\frac{1}{2}(\partial_x -i \partial_y)$
and $a_{\xi}\equiv\frac{1}{2}(a_x-ia_y)$. With this expression one
may consider the phase picked up by the quasi-particle when it
is moved adiabatically around a closed loop, which is given
by $I_a \equiv \oint\mathrm{d}{\bf x}\cdot {\bf a}$. Comparing
this with the phase a moving particle acquires in the system,
$I_A \equiv \oint \mathrm{d}{\bf x}\cdot{\bf A}$,
where ${\bf A}$ is the external gauge potential, one is able to deduce
the fractional charge and fractional statistics of the quasi-particles.
If both loop integrals are equal, i.e. $\eta\equiv I_a/I_A=1$, the
quasi-particle behaves like a normal particle, while the mismatch
by a fractional factor, i.e. $\eta=1/p$ with $p$ an integer, allows to
interpret the quasi-particle as a ``fractional''particle.
The difficulty of this analysis lies in the calculation of
the normalization factors. For the Laughlin state one usually
circumvents this by applying the Plasma analogy~\cite{Laughlin:1983,arovas}
to determine the normalization factor of the corresponding
quasi-holes and quasi-particles, avoiding the explicit calculation.
It is found that $p=m$, i.e. the fractional behavior of the
excitation follows from the fractional filling of the lowest
Landau level. The plasma analogy, however, is not applicable
to all the relevant states, which exhibit such anyonic excitations,
but are different from the Laughlin. Calculating the normalization
factors by direct integration is much too complicated, even for
systems of only a few particles. However, by transforming the
corresponding quasi-hole or quasi-particle state into the
many-body basis by means of our code, the normalization factors are
obtained by simply taking the scalar product of the decomposition vector.
As an example, we calculate $\eta_{qh}$ and $\eta_{qp}$ for
quasi-holes and quasi-particles in small Laughlin systems.
First, we notice that the polynomial in Eqs.~(\ref{lqp}) and
(\ref{lqh}) contains terms of different order in $z$. It is
therefore necessary to separate the $N$ contributions with
fixed angular momentum, and then apply the decomposition explained
in Sect. \ref{sec:deco} to each of them. Then we can write the
normalization factor as a polynomial in $\xi$ and $\xi^*$. For
instance, a the wavefunction of a quasi-hole in the $m=2$ Laughlin
state of $N=5$ bosons, is found to have the following normalization factor:
\begin{eqnarray}
C \propto 1 + 0.477 |\xi|^2 + 0.117 |\xi|^4 + 0.0211 |\xi|^6 + 0.00334 |\xi|^8
+ 0.000668 |\xi|^{10},
\end{eqnarray}
which, even in high orders of $\xi$, agrees reasonably well with
the prediction by the plasma analogy, according to which
$C \propto \exp(\frac{1}{m}|\xi|^2)$.
However, from Table~\ref{fc} showing the results for different
numbers of particles, we find that, especially for the quasi-particle
excitation, significant deviations occur for very small systems
$(N=4)$. For $N=6$, the mismatch is almost cured, and the agreement
with the plasma analogy prediction is better than 3\% for both
quasi-holes and quasi-particles.
\begin{table}
\begin{center}
\begin{tabular}{|l|c|c|}
\hline
$N$ & $\eta_{qh}$ & $\eta_{qp}$ \\
\hline
4 & 0.473 & 0.354 \\
5 & 0.477 & 0.438 \\
6 & 0.487 & 0.501 \\
\hline
\end{tabular}
\caption{\label{fc} Fractional charge of quasi-holes and quasi-particles
in a bosonic Laughlin system at filling $\nu=1/2$ for different number
of particles $N$.}
\end{center}
\end{table}
\section{Conclusion}
We have presented a code which, by means of the symbolic
package Mathematica, analytically decomposes relevant
analytical, strongly-correlated many-body states into
the many-body basis built up by single-particle angular
momentum eigenstates. This basis is commonly used to describe
2D quantum systems subjected to gauge fields, and thus the
described decomposition is a useful tool for calculating
the overlap of different states with famous test states
like the Laughlin state, the Pfaffian state and many others.
It also allows for studying the angular-momentum distribution
of strongly correlated states, which can be related to the
ground-state density of the system, or normalization constants
of the states. The latter has been shown to provide insights
into the fractional character of quasiparticle excitations.
\section*{Acknowledgments}
The authors thank Frank Tabakin and Nuria Barber\'an for a
careful reading of the manuscript and useful suggestions.
B.~J.-D. is supported by a Grup Consolidat SGR 21-2009-2013.
This project is partially supported by Grants No. FIS2008-01661
(Spain), and No. 2009SGR1289 from Generalitat de Catalunya.
We acknowledge financial support from the Spanish MEC project
TOQATA and the ERC Advanced Grant QUAGATUA.
|
\section{Introduction}
Entanglement is not only an essential feature of quantum mechanics,
which distinguishes the quantum from classical world, but also a
great resource in the fields of quantum information and quantum
computation [1-3]. Particularly, entangled qubits prepared in the
pure maximally entangled states, \emph{i.e}., the Bell states, are
required by many quantum information processes [4,5]. However, in
the real world, a pure state in a quantum system will always evolves
to a mixed one due to the unavoidable interactions with the
environment. Thus, for practical purpose, applications of quantum
information processes utilizing the mixed states are under
consideration.
Among the bipartite entangled mixed states, the Bell diagonal state
(BDS) plays an important role in quantum information processing. It
is widely used in the processes of quantum teleportation \cite{m1},
quantum entanglement purification [7,8], quantum key distribution
\cite{m4}, etc. Moreover, the BDS is a simple but significant
example in studying the nonclassical correlation of a quantum mixed
state [10-12], since there always exists a local transformation
which can transform the given mixed state to a corresponding Bell
diagonal form \cite{Cen}. Therefore, identification of an unknown
BDS is of great importance. Conventionally, the identification is
achieved by the so-called state tomography technique [14-16], which
performs the projection measurements on the unknown state directly,
and repeats the measurements on many copies of state. It is a
drawback that after the projection measurements the state to be
measured will collapse to one of the measurement basis. Thus the
original state will be destroyed and become useless.
Recently, schemes for quantum non-demolition measurement are
proposed to detect unknown quantum state [17-19], by which the
detected state will not be destroyed after the measurement. In this
paper we present an alternative scheme for identifying an unknown
BDS. In our scheme, we do not perform the projective measurements on
the BDS directly, but on the probe qubits. According to the
measurement outcomes of the probe qubits, we can acquire all the
information of the unknown BDS. The distinguished advantage of our
scheme is that the BDS is not destroyed by the measurements, since
the evolved BDS plus the probe states will collapse to the original
BDS after the measurement on the probe qubits. Contrast to the
identification scheme with state tomography technique, which is
achieved by sacrificing numerous copies of the unknown state, our
scheme is economic and the resulting BDS is recyclable. The paper is
organized as follows. In Sec. II, we explicitly demonstrate our
scheme in theory. In Sec. III, we discuss the experimental
realization of our scheme in the framework of cavity quantum
electrodynamics (QED). The conclusion is drawn finally.
\section{Scheme for identification of unknown Bell diagonal state}
In this section, we will illustrate our scheme explicitly. The BDS
is a mixture of the well-known Bell states, it is parameterized by
four real numbers $c_1,c_2,c_3,c_4\in[0,1]$, which satisfy the
normalizing condition $\sum_ic_i=1$. Generally, a BDS can be
described as $\rho_{12}=\sum_ic_i|\Psi_i\rangle\langle\Psi_i|$,
where
$|\Psi_{1}\rangle=(|1\rangle_1|0\rangle_2+|0\rangle_1|1\rangle_2)/\sqrt{2}$,
$|\Psi_{2}\rangle=(|1\rangle_1|0\rangle_2-|0\rangle_1|1\rangle_2)/\sqrt{2}$,
$|\Psi_{3}\rangle=(|1\rangle_1|1\rangle_2+|0\rangle_1|0\rangle_2)/\sqrt{2}$,
and
$|\Psi_{4}\rangle=(|1\rangle_1|1\rangle_2-|0\rangle_1|0\rangle_2)/\sqrt{2}$
are the Bell states, the subscript outside the ket denotes the label
of qubit. Here $|0\rangle=[1,0]^\mathrm{T}$ and
$|1\rangle=[0,1]^\mathrm{T}$ are the computational basis, the
superscript $\mathrm{T}$ denotes the matrix transpose.
In order to identify an unknown BDS, we need a probe qubit (labeled
3). The probe qubit interacts with the BDS and extracts the
information from the BDS. We accomplish our scheme by three steps,
in each step we can build an equation between the unknown parameters
and the observable of the probe qubit. With the three steps done, we
will obtain three independent linear equations, from which we can
calculate the parameters. The three steps of our scheme are
demonstrated in the following paragraphs.
\emph{Step 1.} Assume that the probe qubit is in state
$|0\rangle_3$, thus the initial state of the joint system consists
of the BDS and the probe qubit is given by
$\rho^0=\rho_{12}\otimes|0\rangle_3\langle0|$. We perform an unitary
operation $U^1$, given as follows, on the joint three-qubit state,
\begin{equation}
U^1=\frac{1}{2}\left(
\begin{array}{rrrrrrrr}
1 & 0 & 0 & -i & 0 & -i & -1 & 0 \\
0 & 1 & -i & 0 & -i & 0 & 0 & -1 \\
0 & -i & 1 & 0 & -1 & 0 & 0 & -i \\
-i & 0 & 0 & 1 & 0 & -1 & -i & 0 \\
0 & -i & -1 & 0 & 1 & 0 & 0 & -i \\
-i & 0 & 0 & -1 & 0 & 1 & -i & 0 \\
-1 & 0 & 0 & -i & 0 & -i & 1 & 0 \\
0 & -1 & -i & 0 & -i & 0 & 0 & 1 \\
\end{array}
\right)
\label{as}
\end{equation}
As a result the state of the joint system evolves to
$\rho^1=U^1\rho^0U^{1\dagger}$. We can obtain the reduced density
matrix of the probe qubit by tracing over qubits 1 and 2 as follows,
\begin{equation}
\rho_3^1=\left(
\begin{array}{cc}
c_1+c_3 & 0 \\
0 & c_2+c_4 \\
\end{array}
\right).
\end{equation}
One can find that the information of the BDS is carried by the probe
qubit. Performing a $\sigma^z$ measurement on the probe qubit, we
can obtain the following equation between the unknown parameters and
the observable of the probe qubit,
\begin{equation}
M^1=\mathrm{Tr}(\sigma^z_3\rho^1_3)=c_1+c_3-c_2-c_4.
\end{equation}
Tracing over the probe qubit we can obtain the reduced density
matrix of the resulting BDS as follows,
\begin{equation}
\rho^1_{12}=\frac{1}{2}\left(
\begin{array}{cccc}
c_1+c_4 & 0 & 0 & c_1-c_4 \\
0 & c_3+c_2 & c_3-c_2 & 0 \\
0 & c_3-c_2 & c_3+c_2 & 0 \\
c_1-c_4 & 0 & 0 & c_1+c_4 \\
\end{array}
\right).
\end{equation}
Note that underwent the $U^1$ operation, the resulting BDS becomes
different from the original state because the $|\Psi_1\rangle$ and
$|\Psi_3\rangle$ ingredients have exchanged mutually. Fortunately,
we can recover it to the original form by repeating the
above-mentioned process once more with a new probe qubit to exchange
$|\Psi_1\rangle$ and $|\Psi_3\rangle$ ingredients again. It is
interesting that at the end of the recovering process, we can obtain
the same reduced density matrix of the new probe qubit as shown in
Eq. (2) and consequently yield equation Eq. (3) from the new
resulting probe qubit.
\emph{Step 2}. In this step, the probe qubit is also initialized in
$|0\rangle_3$, thus the joint system is in state
$\rho_{12}\otimes|0\rangle_3\langle 0|$. We perform an unitary
operation named $U^2$ on the joint three-qubit state, $U^2$ has the
following form,
\begin{equation}
U^2=\frac{1}{2}\left(
\begin{array}{rrrrrrrr}
1 & 0 & 0 & i & 0 & i & 1 & 0 \\
0 & 1 & -i & 0 & -i & 0 & 0 & 1 \\
0 & -i & 1 & 0 & -1 & 0 & 0 & i \\
i & 0 & 0 & 1 & 0 & -1 & -i & 0 \\
0 & -i & -1 & 0 & 1 & 0 & 0 & i \\
i & 0 & 0 & -1 & 0 & 1 & -i & 0 \\
1 & 0 & 0 & -i & 0 & -i & 1 & 0 \\
0 & 1 & i & 0 & i & 0 & 0 & 1 \\
\end{array}
\right).
\end{equation}
After $U^2$ operation, the probe qubit evolves to the following
form,
\begin{equation}
\rho^2_3=\left(
\begin{array}{cc}
c_1+c_4 & 0 \\
0 & c_2+c_3 \\
\end{array}
\right).
\end{equation}
Performing a $\sigma^z$ measurement on the probe qubit, we can
obtain the following equation,
\begin{equation}
M^2=\mathrm{Tr}(\sigma^z_3\rho_3^2)=c_1+c_4-c_2-c_3.
\end{equation}
The resulting BDS underwent the $U^2$ operation is given as follows,
\begin{equation}
\rho^2_{12}=\frac{1}{2}\left(
\begin{array}{cccc}
c_3+c_1 & 0 & 0 & c_3-c_1 \\
0 & c_4+c_2 & c_4-c_2 & 0 \\
0 & c_4-c_2 & c_4+c_2 & 0 \\
c_3-c_1 & 0 & 0 & c_3+c_1 \\
\end{array}
\right),
\end{equation}
Similar to step 1, we can transform the resulting BDS to the
original form by performing $U^2$ on the joint system which is
composed of the resulting BDS and a new probe qubit. Again the new
resulting probe qubit ensemble will carry the information of the
unknown parameters.
\emph{Step 3.} We perform an unitary operation $U^3$ on the joint
BDS and probe qubit system, $U^3$ is given as follows,
\begin{equation}
U^3=\left(
\begin{array}{cccccccc}
-i & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\
0 & i & 0 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 1 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 & 1 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 & 0 & i & 0 \\
0 & 0 & 0 & 0 & 0 & 0 & 0 & -i \\
\end{array}
\right).
\end{equation}
Different from the previous two steps, here the probe qubit is
initialized in the superposition state
$(|0\rangle_3+|1\rangle_3)/\sqrt{2}$. After $U^3$ operation, the
probe qubit evolves to the following form,
\begin{equation}
\rho^3_3=\frac{1}{2}\left(
\begin{array}{cc}
1 & c_1+c_2-c_3-c_4 \\
c_1+c_2-c_3-c_4 & 1 \\
\end{array}
\right).
\end{equation}
Now, performing a $\sigma^x$ measurement on the probe qubit, we can
obtain the following equation,
\begin{equation}
M^3=\mathrm{Tr}(\sigma^x_3\rho_3^3)=c_1+c_2-c_3-c_4.
\end{equation}
In the meantime, the resulting BDS has the following form,
\begin{equation}
\rho^3_{12}=\frac{1}{2}\left(
\begin{array}{cccc}
c_3+c_4 & 0 & 0 & c_4-c_3 \\
0 & c_1+c_2 & c_1-c_2 & 0 \\
0 & c_1-c_2 & c_1+c_2 & 0 \\
c_4-c_3 & 0 & 0 & c_3+c_4 \\
\end{array}
\right).
\end{equation}
To recover this BDS back to the original state, we repeat $U^3$ on
the joint system composed of the resulting BDS and a new probe qubit
ensemble in state $(|0\rangle_3+|1\rangle_3)/\sqrt{2}$. Obviously,
the new resulting probe qubit ensemble will also carry the
information of the unknown BDS.
Combining equations (3), (7), and (11), and taking into account the
normalizing condition $\sum_ic_i=1$, we can work out the parameters
as:
\begin{equation}
c_1=\frac{M^1+M^2+M^3+1}{4},
\end{equation}
\begin{equation}
c_2=\frac{1-M^1-M^2+M^3}{4},
\end{equation}
\begin{equation}
c_3=\frac{1+M^1-M^2-M^3}{4},
\end{equation}
\begin{equation}
c_4=\frac{1-M^1+M^2-M^3}{4}.
\end{equation}
Now we have succeeded in identifying an unknown BDS with the help of
the probe qubit ensembles. Notably, due to the recovering process in
each step, the final BDS is the same as the initial state.
It is necessary give a discussion on the principles of our scheme.
We emphasize that our scheme is based on the ensemble viewpoint, by
which the BDS can be considered as a mixture of the four Bell
states. The mixing proportion of each Bell state is denoted by the
parameter $c_i$. Each pair of qubits 1 and 2 fetching from the BDS
ensemble will be randomly in one of the four Bell states. Without
loss of generality, we take step 1 as an example to show how the
probe qubit can extract information from the BDS ensemble. The
expression of $U^1$ can be rewritten as
$U^1=|\psi_{2,0}\rangle\langle\psi_{2,0}|+|\psi_{2,1}\rangle\langle\psi_{2,1}|+|\psi_{4,0}\rangle\langle\psi_{4,0}|+|\psi_{4,1}\rangle\langle\psi_{4,1}|-i|\psi_{1,0}\rangle\langle\psi_{3,1}|-i|\psi_{1,1}\rangle\langle\psi_{3,0}|-i|\psi_{3,0}\rangle\langle\psi_{1,1}|-i|\psi_{3,1}\rangle\langle\psi_{1,0}|
$, where $|\psi_{i,j}\rangle=|\Psi_i\rangle\otimes|j\rangle_3$. One
can find that if the state of qubits 1 and 2 is $|\Psi_2\rangle$ or
$|\Psi_4\rangle$, it will remain unchanged and the probe qubit 3
will stay in $|0\rangle_3$; if the state of qubits 1 and 2 is
$|\Psi_1\rangle$ ($|\Psi_3\rangle$), it will change to
$|\Psi_3\rangle$ ($|\Psi_1\rangle$) and flip the probe qubit state
from $|0\rangle_3$ to $|1\rangle_3$. Repeatedly perform $U^1$ on the
joint three-qubit state by fetching new qubits from the BDS and the
probe qubit ensemble, the resulting probe qubit ensemble will end in
a mixed state ensemble which reveals the information of $c_1$ and
$c_3$ through the appearance probability of $|1\rangle_3$. To
transform the resulting BDS back to the original form, we only need
to repeat this process once more to make a simple exchange of
$|\Psi_1\rangle$ and $|\Psi_3\rangle$. In steps 2 and 3, our scheme
works similarly. As a consequence, the information of the unknown
BDS is transferred to the probe ensembles. It is interesting that
the resulting probe ensembles produced by the recover process are
also useful.
Let us look back to the expressions of $U^1$, $U^2$, and $U^3$.
These operators are essentially tripartite manipulations on qubits,
and they can be formally factorized as $U^i=(U^i_{13}\otimes
I_2)(I_1\otimes U^i_{23}) (i=1,2,3)$, where $I_1$ and $I_2$ are the
identity operators of subsystems 1 and 2, respectively. The
bipartite operations $U^i_{13}$ and $U^i_{23}$ are given as follows,
\begin{equation}
U^{1}_{13}=U^1_{23}=\frac{1}{\sqrt{2}}\left(
\begin{array}{cccc}
1 & 0 & 0 & -i \\
0 & 1 & -i & 0 \\
0 & -i & 1 & 0 \\
-i & 0 & 0 & 1 \\
\end{array}
\right),
\end{equation}
\begin{equation}
U^2_{13}=U^2_{23}=\frac{1}{\sqrt{2}}\left(
\begin{array}{cccc}
1 & 0 & 0 & i \\
0 & 1 & -i & 0 \\
0 & -i & 1 & 0 \\
i & 0 & 0 & 1 \\
\end{array}
\right),
\end{equation}
\begin{equation}
U^3_{13}=U^3_{23}=\frac{1}{\sqrt{2}}\left(
\begin{array}{cccc}
1-i & 0 & 0 & 0 \\
0 & 1+i & 0 & 0 \\
0 & 0 & 1+i & 0 \\
0 & 0 & 0 & 1-i \\
\end{array}
\right).
\end{equation}
Based on these factorizations we can accomplish each step by
sequentially performing bipartite manipulation $U^i_{13}$ on qubits
1 and 3, and $U^i_{23}$ on qubits 2 and 3. That is to say we can
perform only bipartite manipulations in the whole processing of our
scheme, instead of tripartite manipulations which is difficult to
realize in experiments. The procedures are given as follows. Suppose
that qubit 1 together with qubit 3 locates at place A, and qubit 3
locates at place B. In each step, we first perform operations
$U^{i}_{13}$ on qubits 1 and 3, next send qubit 3 to place B, and
then perform operations $U^i_{23}$ on qubits 2 and 3. Finally, we
perform measurements on the probe qubit.
\section{Identification of BDS in experimental scenario}
In this section we will discuss experimental realization of our
scheme in the framework of cavity QED system. This experimental
scenario is based on the case that two qubits are separated into
different places, since two-qubit manipulation is more feasible than
the three-qubit manipulation. The schematic illustration is shown in
Fig. \ref{1}(a). Assume that the BDS ensemble is shared by two
participators, each of whom has an optical cavity A and B,
respectively. The particles in the ensemble are considered to be
three-level atoms with two ground states $|a\rangle$ and $|b\rangle$
and an excited state $|e\rangle$, see Fig. \ref{1}(b). The
long-lived levels $|a\rangle$ and $|b\rangle$ represent states
$|0\rangle$ and $|1\rangle$, respectively. The probe atoms are
identical to those in the BDS ensemble. The cavity couples the
atomic transitions $|a\rangle\leftrightarrow|e\rangle$ and
$|b\rangle\leftrightarrow|e\rangle$ with the coupling strength $g_a$
and $g_b$, respectively. Additionally, two external driving lasers
couple the transitions $|a\rangle\leftrightarrow|e\rangle$ and
$|b\rangle\leftrightarrow|e\rangle$ with the Rabi frequencies
$\Omega_a$ and $\Omega_b$, respectively. All the atoms couple to the
cavities via the same mechanism.
\begin{figure}[tbp]
\includegraphics[width=1\columnwidth]{Fig1.eps}
\caption{(Color online) (a) Schematic illustration for
identification of unknown BDS based on the cavity QED system. The
probe atom 3 first interacts with atom 1 in cavity A and then is
sent to cavity B to interacts with atom 2. Each atom in the cavity
is driven by two classical lasers. (b) Atomic levels and
transitions.} \label{1}
\end{figure}
To start the scheme, we pick up a pair of entangled atoms from the
BDS ensemble and put them into the corresponding cavities. The probe
atom 3 is sent into cavity A firstly. The Hamiltonian in cavity A
can be written as follows,
\begin{eqnarray}
H_A&=&\sum_{j=1,3}\omega_e|e\rangle_j\langle
e|+\omega_{ab}|b\rangle_j\langle
b|+\omega_ca^{\dagger}a\cr\cr&&+(\Omega_ae^{-i\omega_at}+g_aa)|e\rangle_j\langle
a|\cr\cr&&+(\Omega_be^{-i\omega_bt}+g_ba)|e\rangle_j\langle b|,
\end{eqnarray}
where $\omega_e$ is the energy of $|e\rangle$ while $\omega_{ab}$ is
the energy of $|b\rangle$, $\omega_a$ ($\omega_b$) is the frequency
of the driver laser with Rabi frequency $\Omega_a$ ($\Omega_b$), and
$a$ is the annihilation operator of the cavity.
By setting $\delta_1=\omega_{ab}-(\omega_a-\omega_b)/2$, we switch
to an interaction picture with respect to
$H_0=\sum_{j=1,3}\omega_e|e\rangle_j\langle
e|+(\omega_{ab}-\delta_1)|b\rangle_j\langle
b|+\omega_ca^{\dagger}a$. Under the large detuning condition
$|\delta_a|, |\delta_b|, |\Delta_a|, |\Delta_b|\gg|g_a|, |g_b|,
|\Omega_a|, |\Omega_b|$, where $\delta_a=\omega_e-\omega_c$,
$\delta_b=\omega_e-\omega_c-\omega_{ab}+\delta_1$,
$\Delta_a=\omega_e-\omega_a$,
$\Delta_b=\omega_e-\omega_b-\omega_{ab}+\delta_1$, we can
adiabatically eliminated the excited state $|e\rangle_j$ [20-23]. If
there are no photons in the cavity and the detunings satisfy
$|\delta_a-\Delta_b+g_a^2/\delta_a|,
|\delta_b-\Delta_a+g_a^2/\delta_a|\gg|\Omega_ag_b/\Delta_a|,
|\Omega_bg_a/\Delta_b|$ (we have assumed
$g_a^2/\delta_a=g_b^2/\delta_b$), considering the subspace without
real photons, we deduce the effective Hamiltonian as
\begin{equation}
H_{\mathrm{eff}}=\frac{2(J_1\sigma_1^++J_2\sigma_1^-)(J_1\sigma_3^-+J_2\sigma_3^+)}{\delta_a-\Delta_b+g_a^2/\delta_a}+B(\sigma_1^z+\sigma_3^z),
\end{equation}
where $\sigma_i^+=|b\rangle_i\langle a|$ and
$\sigma_i^-=|a\rangle_i\langle b|$, the coefficients are given as
\begin{equation}
J_1=\frac{g_a\Omega_b}{2}(\frac{1}{\delta_a}+\frac{1}{\Delta_b}),
\end{equation}
\begin{equation}
J_2=\frac{g_b\Omega_a}{2}(\frac{1}{\delta_b}+\frac{1}{\Delta_a}),
\end{equation}
\begin{eqnarray}
B&=&\frac{\Omega_a^2\Omega_b^2/4}{\delta_a-\delta_b}(\frac{1}{\Delta_a}+\frac{1}{\Delta_b})^2+\frac{\Omega_b^2g_b^2/4}{\delta_b-\Delta_a+g_a^2/\delta_a}(\frac{1}{\Delta_b}+\frac{1}{\delta_b})^2
\cr\cr&&-\frac{\Omega_a^2g_a^2/4}{\delta_a-\Delta_a+g_a^2/\delta_a}(\frac{1}{\Delta_a}+\frac{1}{\delta_a})^2
\cr\cr&&+\frac{1}{2}(\frac{\Omega_b^2}{\Delta_b}-\frac{\Omega_a^2}{\Delta_a}+\delta_1).
\end{eqnarray}
The effective magnetic field $B$ can be tuned to be very close to
zero by varying $\delta_1$. For
$\Omega_b=g_b\Omega_a(\frac{1}{\Delta_a}+\frac{1}{\delta_b})/[g_a(\frac{1}{\delta_a}+\frac{1}{\Delta_b})]$,
we can get the final effective Hamiltonian as follows,
\begin{equation}
H_{xx}=\lambda_x\sigma_1^x\sigma_3^x.
\end{equation}
where $\lambda_x=2J_1^2/(\delta_a-\Delta_b+g_a^2/\delta_a)$. It is
obvious to see that the unitary time-evolution operator
$e^{-iH_{xx}t}$ is in accordance with the unitary operator
$U_{13}^1$ at $t=\frac{(2n+1)\pi}{4\lambda_x},n=0,1,2...$, thus we
realize the unitary operation $U_{13}$. To realize the operation
$U_{23}^2$, we send the probe atom to cavity B, and drive atoms 2
and 3 with the same lasers as done in cavity A.
If we select the Rabi frequency as
$\Omega_b=-g_b\Omega_a(\frac{1}{\Delta_a}+\frac{1}{\delta_b})/[g_a(\frac{1}{\delta_a}+\frac{1}{\Delta_b})]$,
we can obtain
\begin{equation}
H_{yy}=\lambda_y\sigma_1^y\sigma_3^y.
\end{equation}
where $\lambda_y=2J_1^2/(\delta_a-\Delta_b+g_a^2/\delta_a)$. At time
$t=\frac{(2n+1)\pi}{4\lambda_y},n=0,1,2...$, the time-evolution
unitary operator $e^{-iH_{yy}t}$ coincides with the operator
$U^2_{13}$. Then sent the probe atom into cavity B, we can realize
the unitary operator $U^2_{23}$ by controlling the interaction time.
In order to realize the operations $U^3_{13}$ and $U^3_{23}$, we
choose the laser frequencies as $\omega_a=\omega_b=\omega$. We
switch to an interaction picture with respect to
$H_0=\sum_{j=1,3}\omega_e|e\rangle_j\langle
e|+(\omega_{ab}-\tilde{\delta}_1)|b\rangle_j\langle
b|+\omega_ca^{\dagger}a$, where the detuning $\tilde{\delta}_1$ is
introduced to tune the effective magnetic field. Under the large
detuning condition we can adiabatically eliminated the excited
states. If there are no photons in the cavity and the detunings
satisfy $|\delta_a-\Delta_a+g_a^2/\delta_a|,
|\delta_b-\Delta_b+g_a^2/\delta_a|\gg|\Omega_ag_a/\Delta_a|,
|\Omega_bg_b/\Delta_b|$ (assuming $g_a^2/\delta_a=g_b^2/\delta_b$),
considering a subspace with no real photons we can obtain the
following effective Hamiltonian,
\begin{eqnarray}
H_{\mathrm{eff}}&=&\frac{2(\tilde{J}_1|a\rangle_1\langle
a|+\tilde{J}_2|b\rangle_1\langle b|)(\tilde{J}_1|a\rangle_2\langle
a|+\tilde{J}_2|b\rangle_2\langle
b|)}{\delta_a-\Delta_a+g^2_a/\delta_a}\cr\cr&&+\tilde{B}(\sigma_1^z+\sigma^z_3),
\end{eqnarray}
where $\sigma^z_i=|b\rangle_i\langle b|-|a\rangle_i\langle a|$, and
the coefficients are given as follows,
\begin{equation}
\tilde{J}_1=\frac{\Omega_ag_a}{2}(\frac{1}{\Delta_a}+\frac{1}{\delta_a}),
\end{equation}
\begin{equation}
\tilde{J}_2=\frac{\Omega_bg_b}{2}(\frac{1}{\Delta_b}+\frac{1}{\delta_b}),
\end{equation}
\begin{eqnarray}
\tilde{B}&=&\frac{1}{2}(\frac{\Omega^2_b}{\Delta_b}+\tilde{\delta}_1-\frac{\Omega^2_a}{\Delta_a})+\frac{1}{2}[\frac{\Omega_b^2g_a^2/4}{\delta_a-\Delta_b+g_a^2/\delta_a}(\frac{1}{\delta_a}+\frac{1}{\Delta_b})^2\cr\cr&&-\frac{\Omega_a^2g_b^2/4}{\delta_b-\Delta_a+g_a^2/\delta_a}(\frac{1}{\delta_b}+\frac{1}{\Delta_a})^2]\cr\cr&&+\frac{\Omega_a^2\Omega_b^2/4}{\Delta_a-\Delta_b}(\frac{1}{\Delta_a}+\frac{1}{\Delta_b})^2.
\end{eqnarray}
The effective magnetic field can be tuned to zero by varying
$\tilde{\delta}_1$. For
$\Omega_b=-\Omega_ag_a(\frac{1}{\Delta_a}+\frac{1}{\delta_a})/[g_b(\frac{1}{\Delta_b}+\frac{1}{\delta_b})]$,
we can obtain the following effective Hamiltonian,
\begin{equation}
H_{zz}=\lambda_z\sigma_1^z\sigma_3^z,
\end{equation}
where $\lambda_z=2\tilde{J}_1/(\delta_a-\Delta_a+g_a^2/\delta_a)$.
It is obvious to see that the time-evolution unitary operator
$e^{-iH_{zz}t}$ coincides with $U_{13}^3$ at time points
$t=\frac{(2n+1)\pi}{4\lambda_z}$ $(n=0,1,2,...)$. Thus we have
realized the operation $U^3_{13}$ in cavity A, in the same way we
can realize the operation $U^3_{23}$ in cavity B by sending the
probe atom into cavity B.
To confirm the validity of our approximation, we numerically
simulate the dynamics generated by the full Hamiltonian and compare
it to the the dynamics generated by the effective Hamiltonian. In
Fig. 2(a), we have plotted the time evolution of
$\langle\sigma_3^z\rangle$. The numerical results show that the
performance of $\langle\sigma_3^z\rangle$ under the full Hamiltonian
and that under $H_{xx}$ agree with each other reasonably well.
Similar agreement can also be seen in Fig. 2(b) and (c). Therefore,
our effective model is valid.
\begin{figure}[tbp]
\includegraphics[width=1\columnwidth]{Fig2.eps}
\caption{(Color online) Time evolutions of atom 3 calculated using
full Hamiltonian (dashed blue line) and effective Hamiltonians
(solid red line). The effective Hamiltonians for (a), (b), and (c)
are $H_{xx}$, $H_{yy}$, and $H_{zz}$, respectively. The initial
states are chosen as $0.5(|a\rangle_1\langle a|+|b\rangle_1\langle
b|)\otimes|a\rangle_3\langle a|$ for (a) and (b), and
$0.25(|a\rangle_1\langle a|+|b\rangle_1\langle
b|)\otimes(|a\rangle+|b\rangle)_3(\langle a|+\langle b|)$ for (c).
The parameters are chosen as $g_a=1$, $\Omega_a=5g_a$,
$\delta_a=102g_a$, $\delta_b=122g_a$, $\Delta_a=120g_a$, and
$\Delta_b=100g_a$ for (a) and (b); $g_a=1$, $\Omega_a=5g_a$,
$\delta_a=102g_a$, $\delta_b=122g_a$, $\Delta_a=100g_a$, and
$\Delta_b=120g_a$ for (c). The inset shows the behaviors of the two
curves in detail with time $t\in[500,510]$ (in units of $1/g_a$).}
\label{2}
\end{figure}
So far, we have realized all the unitary operations described by
Eqs. (17)-(19) in the cavity QED system. Since the qubits are
encoded in the ground atomic states and there is no real photons in
the cavity, this experimental scenario is robust against the
dissipative effects.
Here we give a brief discussion on the experimental feasibility of
the presented scheme. For an experimental implementation, the
effective coupling strengths $\lambda_x$, $\lambda_y$, and
$\lambda_z$ should be much larger than the cavity leaky rate
$\Gamma_C$ and the atomic spontaneous rate $\Gamma_E$. This
requirements can be satisfied in microcavities which have a small
volume and thus a high quality factor. Suitable candidates for the
present proposal are, for example, the microtoroidal cavities which
has cooperativity factor $g^2/(\Gamma_C\Gamma_E)\sim10^7$ and the
ratio $g/\Gamma_E\sim10^3$ [23,24], where $g$ is defined by
$g=\mathrm{max}(g_a,g_b)$. Thus our scheme is feasible with current
available systems.
\section{Conclusion}
In conclusion, we have presented a scheme for nondestructive
identifying unknown BDS by measuring the probe qubits. This scheme
is implemented in three steps. In each step we can build an equation
between the unknown coefficients of the BDS and the observable of
the probe qubit. Combining the three equations we can calculate the
parameters. Moreover, at the end of each step, the BDS ensemble
remains in the initial state, therefore it is not polluted by the
identification processing. We also consider the experimental
realization of the scheme in the cavity QED system. By selecting
appropriate Rabi frequencies of the driving lasers, we can realize
the corresponding unitary operations, respectively. Our scheme is
feasible with the current techniques.
\section{Acknowledgement}
This work was supported by the National Natural Science Foundation
of China, under Grants No. 10805007 and No. 10875020, and the
Doctoral Startup Foundation of Liaoning Province.
|
\section{Introduction}
Multimessenger astronomy with massive black hole (MBH) binaries (MBHBs) has been discussed frequently in recent literature
(e.g. \cite{HolzHughes2005,Armitage:2005,Kocsis2006,Phinney2009,schnittman2011,roedig11,Sesana11,tanaka11})
as a unique way of fully understanding the mutual interaction of these systems with their environment. The general
picture of MBHBs forming after a galaxy merger has been given in \cite{begelman80}, but observational evidence for
sub--parsec MBHBs is still lacking (see~\cite{colpid2009} for a review). Numerical studies down to parsec
scales \cite{mm01,Escala2005,Dotti07,Dotti2009b} have shown that the two MBHs continue to lose orbital
energy and angular momentum under the large-scale action of gas/star-dynamical friction, and
end up forming a Keplerian binary. At this point, the gaseous and stellar environment
furthers the inspiral by either three-body scattering of individual stars and/or the interaction of
the binary with a circumbinary gaseous disc (e.g. \cite{MerrittReview05,Armitage:2002}). At this late
evolutionary stage, excitation of $e$ was already noticed and studied in both environments
(see, e.g., \cite{Armitage:2005,Sesana2010}).
A non zero source eccentricity has important implications for gravitational wave (GW) observations:
on one hand, eccentricity has an impact on source detection and parameter estimation; on the other
hand, measured eccentricities carry valuable information about the dynamical evolution of the detected
systems. In gas rich environments, the surviving detectable eccentricity in the LISA band
($10^{-4}-10^{-1}$Hz) was addressed in \cite{roedig11},
stating that $e_{\rm LISA}\sim 10^{-3}-10^{-1}$ depending on the MBHB parameters. However, in view of the recent
re--design of the LISA mission, estimates have to be re-adapted accordingly. The New Gravitational Observatory
(ELISA/NGO){\footnote{\url{https://lisa-light.aei.mpg.de/bin/view/}}} will see a shorter portion of the
binary inspiral, with inevitable consequences for parameter estimation, and for residual eccentricity
detectability. Pulsar timing arrays (PTAs) \cite{hobbs10} are expected to detect the first GWs in the low-frequency
band by the end of this decade. Given their frequency window ($10^{-9}-10^{-7}$Hz),
they will be sensitive to very massive ($>10^8\msun$) systems only , whose chirping is still marginal.
This implies at the same time that observed MBHBs might still be coupled to their environment.
We assess here the expected eccentricity distribution for these heavy sources in the PTA band.
We then focus on the possibility of having observable PTA systems that are to some extent still coupled
to either their stellar or gaseous environment, and assess the plausibility of detecting possible EM
counterparts with current and up--coming missions, such as eROSITA \cite{eRosita10}, MAXI \cite{MAXI09}
and ATHENA\footnote{\url{http://www.mpe.mpg.de/athena/home.php?lang=en}}.
The paper is structured as follows: First we describe the mechanisms of eccentricity growth in stellar
environments in Sec.~\ref{sec:stellar} and gaseous discs in Sec.~\ref{sec:gas}. We then calculate
the residual eccentricity distributions in the ELISA/NGO band in Sec.~\ref{sec:ELISA}
and comment on the expected eccentricity population in the PTA band in Sec.~\ref{sec:PTA}. We review
the implications this has on multimessenger signals in Sec.~\ref{sec:multi} and
give our conclusions in Sec.~\ref{sec:conclusions}.
\section{Mechanisms of eccentricity growth:}
In the following we discuss the main physical mechanisms driving the MBHB eccentricity
evolution in the dense astrophysical environment of a post-merger galaxy. We consider
a MBHB with total mass $M=M_1+M_2$ ($M_2<M_1$ are the individual MBH masses),
semimajor axis $a$ and eccentricity $e$. Throughout the manuscript, quantities
without subscripts (such as $M$, $a$, $e$, $E$, $L$) always refer to the MBHB,
whereas subscripts '1' and '2' refer to the individual MBHs.
\subsection{Stellar environment}
\label{sec:stellar}
In dense stellar environments, the MBHB evolution is basically driven by three body
interactions with individual stars \cite{mikkola92,quinlan96,sesana06}. To get a sense of the
general picture, it is useful to consider the interactions between the stars and the binary
one by one, as uncorrelated events, even though secular effects may also play a role (see, e.g., \cite{meiron11}).
At the moment of pairing at sub-pc separations,
the MBHB is expected to be surrounded by a dense stellar distribution of stars.
If $a$ is the MBHB semimajor axis, three body interactions with stars approaching the
binary center of mass within $a$ result in superelastic scattering, usually followed by
the ejection of the stars. These stars carry away energy and angular momentum, causing the
binary to shrink. The efficiency of this mechanism depends on several environmental factors,
such as the details of the stellar distribution and the effectiveness of relaxation mechanisms.
We are here interested in the eccentricity evolution.
Assessing the eccentricity evolution is not straightforward, because it depends on a
combination of energy ($E$) and angular momentum ($L$) exchange during the scattering:
$\Delta{e}\propto c_i\Delta{E} +c_j\Delta{L}$, where the two $c$ coefficients depend
on the property of the MBHB. Let us consider,
for simplicity, a bound star orbiting the MBHB at a given semimajor axis $a_*$. The
orbital energy that can be extracted in the scattering is $\propto a_*^{-1}$, whereas the angular
momentum of the orbit is proportional to $\sqrt{a_*}$. It follows that
$\Delta{e}\propto -c_ia_*^{-1}+c_j\sqrt{a_*}$ \cite{sesana08}.
If $a_*$ is small enough, then the first term dominates
and the binary circularizes; vice versa, if $a_*$ is large enough, then the second term dominates, and
the binary eccentricity grows. This behavior has to be expected, since stars with large $a_*$ have large
angular momentum but just little energy, and the opposite is true for stars with small $a_*$.
It turns out that the transition point is at $a_*\approx a$, i.e., stars within the binary orbit
on average promote circularization, whereas stars outside the binary orbit make the binary grow
eccentric. Heuristically this can be interpreted as follows. Let us consider a MBHB with
$M_2\ll M_1$ and a given non-zero eccentricity. Such a binary spends most of
its period near its apocenter, so in the case $a_* > a$ the probability
of a close star-binary encounter (and subsequent star ejection) is
maximal there. In the unequal mass case, the star is ejected following a series of close
encounters with $M_2$. In such interactions the star gets a small 'kick' and the
velocity of $M_2$ decreases. As this velocity is perpendicular to the
MBH position vector close to the apocenter, the secondary is forced on a more
radial orbit. On the contrary, a star with $a_* < a$ "feels" the secondary
hole when this approaches the primary at a distance $<a$.
At that point, the interaction with the star is unlikely to occur close
to the pericenter of the MBHB orbit (because the time spent by the binary at the pericenter
is very small), and typically also extracts a large radial
component from the velocity of the secondary black hole, hence
causing circularization. Considering the collective effect of the stellar
distribution, the evolution of the orbital eccentricity is therefore
governed by the relative weight of stars inside and outside $a$. The steeper
is the stellar cusp, the less effective is the eccentricity growth.
As a general trend, quasi circular-equal mass MBHBs experience just a mild
eccentricity growth, while systems which are already eccentric
at the moment of pairing, or with mass ratio significantly lower than
1, can evolve up to $e>0.9$ \cite{Sesana2010}. Such a trend has been observed in
several numerical studies \cite{mm01,mms07,mat07,as09}.
The above argument is strictly valid for isotropic stellar distributions.
The evolution of the binary eccentricity can be extremely different
for non isotropic systems. For example, \cite{sgd11}
demonstrated that in a rotating stellar system, the eccentricity
evolution of unequal MBHBs is dramatically affected by the level of
co/counter rotation of the stellar distribution with respect to the
binary, with corotating distributions promoting circularization rather
than eccentricity growth. This is because, as a simple consequence of
angular momentum conservation during the ejection process,
stars counterrotating with the binary tend to extract a lot of angular
momentum from the MBHB, causing the eccentricity growth, whereas
corotating stars do not. Nonetheless, most of the simulations
involving rotating bulges \cite{ber06,Berentzen2009}, or merging systems
\cite{khan11,preto11} find quite eccentric binaries at the
moment of pairing (ranging from $0.4$ to $0.8$), and the subsequent
evolution leads to a general eccentricity growth, in good agreement of
what predicted for an isotropic stellar distribution. This may be
because the binary evolution is mostly driven by loss cone refilling
of unbound stars on almost radial orbits, with negligible initial
angular momentum. On the other hand, \cite{mg11} find milder eccentricity
evolutions, along the lines of what is expected for a rotating stellar
remnant. The eccentricity evolution of MBHBs in stellar environments
is still a largely open issue, but a significant eccentricity seem to be a general
outcome of the models.
\subsection{Gaseous environments}
\label{sec:gas}
The study of the secular evolution of a MBHB immersed in a gaseous disc is essentially similar to
type-II planet migration \cite{GoldreichSari03}. However, when self-gravity of the disc
and the accretion of matter onto the MBHs is included,
the clear predictions of resonances and decay times coming from perturbation theory tend to be
smeared out. In this paragraph, we will sketch the two effects that lead to the excitation and
the saturation of the eccentricity evolution:
\begin{itemize}
\item excitation of eccentricity via asymmetric torques from the circumbinary disc
\item saturation of eccentricity growth due to accumulated gas around the black holes
\end{itemize}
When dynamical friction becomes inefficient, the subsequent MBHB evolution relies on
transfer and redistribution of energy and angular momentum from the binary to the
so--called circumbinary disc.
This usually occurs at sub-parsec binary orbital separations, and goes hand in hand with the
appearance of a so--called \textit{gap}. A \textit{gap} is a hollow, low--density region which
is found to have a radial extent $\delta$ of about twice the semi-major axis $a$ of the binary
(e.g. \cite{MacFadyen2008,Jorge09,roedig11}) which is attributed to the excitation of outer
Lindblad resonances (OLR) with the most prominent one at $\delta_{m=1}=2^{2/3}a \sim 1.59\,a$
\footnote{The excitation of the $m=1$ OLR is suppressed in equal mass binaries, nevertheless
similar gap-sizes are also found for exactly equal mass MBHBs \cite{MacFadyen2008}}.
However, gas will still continue to leak from the inner edge of the disc
to the black holes and form \textit{mini-discs} inside the respective Roche-lobes
\cite{arty96,hayasaki07,roedig11,Sesana11}. As first shown by \cite{Armitage:2005} and confirmed by
\cite{Jorge09}, a binary inside a gap will not remain circular and neither will
the inner edge of the disc (see also \cite{Lubow1991,Papaloizou2001,MacFadyen2008}).\\
As showed in \cite{roedig11} for a MBHB of total mass $M$ and mass-ratio $q= M_2/M_1=1/3$,
the eccentricity of the binary will continue to grow up to about $e\sim 0.6$ while
the eccentricity of the disc stays small, $e_{\rm disc} < 0.15$, even for $e >0.7$.
The reason for this growth of $e$ can be understood by comparing the angular
velocity of the black holes to that of the fluid elements in the circumbinary disc.
The movement of the MBHB inside the disc will induce overdensities close to the inner rim of
the circumbinary disc, moving at lower angular speeds than the MBHs.
If the MBHs are at apocenter, where they are slowest and thus remain the longest,
they will experience a deceleration (negative torque) thus decreasing their
associated angular momenta $L_{1,2}$ and increasing $e$.
Since the disc is at the same time expanding radially to accommodate the larger
maximum MBHB separation, the theoretical limit to this growth depends on the evolution of
$\delta(e,t)$. Equating the average angular velocity of the inner rim of the disc to that of the binary
at apocenter: $\langle \omega_{\rm fluid} \rangle= \omega_{\rm apo}$ yields the implicit relation for a limiting $e$:
\begin{equation}
\frac{(1+e)^3 }{(1-e)}=\delta_{a}(e,t)^3,\,\,\,\,\,\,\,\,\,\delta_{a}=\delta/a
\label{eq:limit1}
\end{equation}
which shifts the question to determining the function $\delta(e,t)$.
Assuming $\delta(e,t)=const=2$, \cite{roedig11} found a limiting eccentricity of $\approx 0.6$. However
we expect $\delta(e,t)$ to generally be a growing
function of $e$ related to the strongest OLRs. A detailed discussion of the OLRs and their evolution will soon
be found in \cite{roediginprep}. In general, if $\delta$ grows with $e$, $L_{\rm fluid} < L_{2}$ is always
true, and $e$ continues to grow unless some other effect comes into play.
When the binary is very eccentric, the Roche-lobes and the mini--discs attached to each individual MBH
overlap at pericenter. Whenever this starts happening the secondary MBH will have to pass through
the mini--disc of the primary and it will be slowed down at pericenter, decreasing $e$. This effect
is thus strongly dependent on the thermodynamics of the gaseous disc and the mass contained within
the mini--disc.
In the remainder of this section, we will elucidate the growth of $e$ with numerical examples of Newtonian
self-gravitating discs around a MBHB with $q=1/3$, using a setup described in \cite{roedig11}, the details
of which will soon be found in \cite{roediginprep}.
The two examples use the same initial data and code, differing only in the thermodynamics
treatment inside the gap. The \textit{dry} run allows for an adiabatic evolution of the internal
energy plus a $\beta$ cooling inside the gap which effectively suppresses gas-inflows, whereas
the \textit{wet} run furthers the inflows by treating the gas isothermally when crossing the
radius of $r_{\rm iso}=1.73\,a$, thus allowing us to differentiate the effects of the gas in the gap.
When plotting the torque density $T$ exerted by the disc onto the MBHs in figure~\ref{fig:torque},
we show the individual torques in the $z$-direction onto each MBH
in the middle and right panels and the sum $T_z= M_1 \, T_{z_1}+ M_2 \, T_{z_2}$ in the left panel.
Since the binary is rotating counterclockwise, the negative portion of the torques always stems
from the upstream region (yellow/red). Comparing the top and the bottom set of panels,
we observe that altering the prescription for the thermodynamics inside the gap alters the torque density
and the dynamical evolution of the entire disc. It is clear that the torque
from the mini--discs are larger for the \textit{wet} run (bottom panels), since the mass inside the mini--discs
is higher and the gas is not allowed to cool (i.e. its internal energy remains constant) after
crossing the threshold radius $r_{\rm iso}$. The torque density onto $M_1$ (middle panel) is more or less
symmetric in the azimuthal direction, whereas $M_2$ clearly experiences larger torques from
the part of the disc that is close to it. What is moreover notable, is the fact that in the left panel, where
the total torque is shown to have a quadrupolar structure,
the contributions from the mini--discs do not have exactly the same orientation as the contributions
from the disc. However, the growth of $e$ is very similar for \textit{wet} and \textit{dry} runs, which leads us to
conclude that the \textit{excitation} of $e$ stems mostly from the circumbinary disc and is robust under
different numerical prescriptions. The precise value of $e_{\rm limit}$, however, depends on the density
and angular momentum of the mini--discs, but is expected to be $e_{\rm limit}\in [0.6,0.8]$.
Further studies of the precise dependence on the thermodynamics are ongoing \cite{roediginprep}.
\begin{figure}
\centering
\includegraphics[width=0.9\linewidth]{plots/adia05_500.ps}\\
\includegraphics[width=0.9\linewidth]{plots/iso05_499.ps}
\caption{Torque density in the $z$-direction for two initially circular MBHB runs reaching
$e\approx0.1$ after $90$ orbits. Top (bottom) panels: \textit{dry} (\textit{wet})
thermodynamic prescription for the cavity. In each line, the three panels are, from left to right:
the total torque density on the binary $T_z= M_1 \, T_{z_1}+ M_2 \, T_{z_2}$,
torque onto the primary MBH $T_{z_1}$, torque onto the secondary MBH $ T_{z_2}$.
Figure made using {\sc SPLASH}\cite{Price2007}}
\label{fig:torque}
\end{figure}
\section{Implications for ELISA/NGO and PTA sources}
\label{sec:implications}
Low-frequency GWs are expected to be rich in astrophysical information if the source parameter
analysis is sufficiently accurate to differentiate between various proposed astrophysical scenarios.
Great hopes lie in a coincident detection of EM counterparts to complement
the GW information with an independent measurement of the source redshift and a connection with
the host galaxy, opening new scenarios for precision cosmology (see, e.g, \cite{HolzHughes2005}).
The proposed ELISA/NGO instrument is designed for $10^3-10^7\msun$ MBHBs
chirping and merging in the $10^{-4}-10^{-1}$ Hz band,
whereas PTAs will be sensitive to GW frequencies in the range $10^{-9}-10^{-6}$Hz with
successful detections depending on the sensitivity of the radio instruments, the number of
available stable millisecond pulsars and the observation time.\\
To guide the development of effective detection pipelines, it is reasonable to review the assumptions
of GW sources emitting purely sinusoidal, monochromatic signals. To estimate the impact of the eccentricity
growth via the mechanisms described above, we need to understand at which binary separation they are effective.
We therefore compare the star/gas driven shrinking timescales to the GW energy loss timescale.
The GW timescale is given by:
\begin{equation}
t_{\rm GW}=7.84\times10^7 \,{\rm yr}\, M_8 q_{\rm s}^{-1} a_3^{4} F(e)^{-1},\,\,\,\,\,\,F(e)=(1-e^2)^{-7/2}\left(1+\frac{73}{24}e^2 +\frac{37}{96}e^4 \right)
\label{tgw}
\end{equation}
where $q_{\rm s}=4q/(1+q^2)$ is the symmetric binary mass
ratio, $M_8=M/10^8\msun$ and $a_3$ is the binary
separation in units of $10^3R_S$ ($R_S=2GM/c^2$).
In gas rich environments, assuming a $\beta$--disc model \cite{ss73} for the circumbinary disc, the
migration timescale is \cite{Sesana11}
\begin{equation}
t_m=2.09\times10^6 \,{\rm yr}\, \alpha_{0.3}^{-1/2}\left(\frac{\dot{m}_{0.3}}{\epsilon_{0.1}}\right)^{-5/8}M_8^{3/4} q_{\rm s}^{3/8}\delta_a(e)^{7/8} a_3^{7/8}.
\label{tm}
\end{equation}
Here we defined the viscosity parameter $\alpha_{0.3}=\alpha/0.3$, the accretion efficiency
$\epsilon_{0.1}=\epsilon/0.1$ and the mass accretion rate $\dot{m}_{0.3}=\dot{m}/0.3$,
where $\dot{m}=\dot{M}/\dot{M}_{\rm Edd}$ is the accretion rate normalized to the Eddington rate.
In a star dominated system, assuming a full loss cone \cite{Sesana2010} the MBHB hardening timescale
is instead given by
\begin{equation}
t_{h}= 2.89\times10^6 \,{\rm yr}\, \sigma_{100} M_8^{-1}\rho_5^{-1} a_3^{-1} H_{15}^{-1},
\end{equation}
where we introduced the stellar velocity dispersion $\sigma_{100}=\sigma/100$ Km s$^{-1}$, the stellar
density at the binary influence radius (see \cite{Sesana2010}, for details) $\rho_5=\rho/10^5\msun$
pc$^{-3}$ and the dimensionless hardening rate $H_{15}=H/15$. Note that the combined $M_8^{-1}$ and
$a_3^{-1}$ scaling implies that $t_h$ is independent on the MBHB mass.
By equating $t_m$ and $t_h$ to $t_{\rm gw}$, one gets the characteristic $a$ at which GW emission takes over
in the binary evolution. It turns out that binaries are still coupled to their environment down to a
separation of few hundred Schwarzschild radii, where the most massive systems fall in the PTA frequency band.
This has two direct consequences: (i)
PTA sources are generally eccentric, with an eccentricity distribution directly determined by the shrinking mechanism,
and (ii) significant residual eccentricity might be present close to coalescence, in the ELISA/NGO band.
In the following we apply selfconsistent models for the evolution of the MBHBs in gaseous and stellar environments
to the expected population detectable by LISA/NGO and PTAs. In gas rich systems, we assume MBHBs to maintain
an eccentricity of $e_{\rm limit}=0.6$ as long as they are coupled to their environment and then to circularize
under the action of GW emission after decoupling. In star dominated systems, we use the self--contained model
for the evolution of the MBHB under the effect of star ejection and GW radiation as outlined in \cite{Sesana2010}.
\subsection{Eccentricity population of ELISA/NGO sources}
\label{sec:ELISA}
\begin{figure}
\includegraphics[width=84.0mm]{plots/fig_ELISA_eccentricity.ps}
\includegraphics[width=84.0mm]{plots/fig_PTA_eccentricity.ps}
\caption{Eccentricity population of MBHBs detectable by ELISA/NGO and PTAs, expected in stellar
and gaseous environments. Left panel: The \textit{solid } histograms represent the efficient
models whereas the \textit{dashed} histograms are for the inefficient models. Right panel:
\textit{solid} histograms include all sources producing timing residuals above 3 ns,
\textit{dashed} histograms include all sources producing residual above 10 ns.}
\label{fig:eccdist}
\end{figure}
In view of the new sensitivity curve of the ELISA/NGO instrument, we re--visit here the
question of the residual detectable eccentricity of a population of MBHBs evolving either
in stellar or gaseous environments. We take the population of merging MBHBs predicted
by a standard MBH cosmic evolution model where MBHs grow by merger and accretion starting
from a population of light seed MBHs, thought to be remnants of the PopIII stars \cite{Madau2001}.
To be specific, we took the model "LE" from \cite{arun09}. We stress here that in this model, MBHBs were
assumed to instantly coalesce after a dynamical friction timescale from the pairing of the
two galaxies hosting the two MBHs: none of the dynamical evolution models is implemented
selfconsistently in deriving the MBHB population. We therefore have a single MBHB population
to which we apply different dynamical evolution models to determine the residual eccentricity.
For each mechanism (gas/star) we consider two scenarios (efficient/inefficient), to give an
idea of the expected eccentricity range. The models are the following
\begin{enumerate}
\item gas-efficient: $\alpha=0.3$, $\dot{m}=1$. The migration timescale is maximized for this high
values of the disc parameters, and the decoupling occurs in the very late stage of the MBHB evolution;
\item gas-inefficient: $\alpha=0.1$, $\dot{m}=0.1$. Decoupling occurs at much larger (factor 3-to-5)
separations, and MBHBs have much more room to circularize;
\item stars-efficient: $p(e_i)\propto e$. Initial eccentricities are taken according to a thermal
probability density function \cite{preto11};
\item stars-inefficient: $e_i=0$. Binaries are initially circular, a condition that minimizes the
effectiveness of eccentricity growth.
\end{enumerate}
Results are shown in the left panel of figure~\ref{fig:eccdist}, where we plot the residual eccentricity of all
sources detected with SNR$>8$, when they enter in the ELISA/NGO band{\footnote{The band entrance
is defined as the frequency where the strain of the source $h_s$ squared equals the one-sided noise
spectral density of the detector $S(f)$.}}. In general, MBHBs evolving in gaseous environments are
expected to retain a larger eccentricity, with a distribution tail possibly extending to $e\sim0.5$.
Residual eccentricities are mostly in the $10^{-4}-10^{-2}$ range, and are generally lower for
star driven systems. Note that the distribution predicted by the star-efficient model just
overlaps to the one predicted by the gas-inefficient one. Therefore, apart from those two extreme cases,
the distributions predicted by the two families of models are generally distinct, especially at the high $e$ tail.
A sufficient source detection statistics should allow discrimination between gas and star driven
dynamics, providing valuable information of the MBHB environment.
\subsection{Eccentricity population of PTA sources}
\label{sec:PTA}
To investigate PTA sources, we instead use MBHB population models from \cite{ks11}. Here we just want to highlight
the fact that PTA sources might be very eccentric, and adequate signal processing algorithms need to be developed.
We therefore use a single representative model for each mechanism. For the gas driven evolution we
use $\alpha=0.3$, $\dot{m}=0.3$; whereas for the star driven model we use $e_i=0.1$ for all the binaries.
In the right panel of figure~\ref{fig:eccdist}, we plot the eccentricity distribution of all sources contributing
to the signal at timing residual level above $10$ns or $3$ns, which is expected to be in the reach of future
PTA campaigns. The distributions are rather similar with more spread and higher extremal eccentricities
for the stellar case. It is not clear, whether the two distributions are discernible; the figures,
however, clearly suggest that a large portion of PTA signals would have fairly high eccentricities
($e>0.1$).
\paragraph{Multimessenger signals.}
\label{sec:multi}
Concentrating on gaseous disc environments, \cite{Sesana11} and \cite{tanaka11} have, in a complementary
way, outlined possible counterparts to resolvable PTA sources, which we will quickly review here. First,
we will look at coupled systems, as done in \cite{Sesana11}, then
conclude with a remark on how eccentricity impacts on decoupled systems as studied in \cite{tanaka11}.\\
If $t_{\nu}$ is the viscous time of the disc, then as long as $t_{GW}>t_{\nu}$, the disc can dissipate
inwards fast enough to follow the MBHB and during this phase the MBHB is expected to be highly eccentric.
In \cite{roedig11}, it has been shown that the strength of a periodicity in the light curve related to the
MBHB orbital frequency depends on the eccentricity of the orbit and the angular momentum
distribution of the inspiralling gas. First, it is important to note that for circular or mildly
eccentric orbits the periodicity in the gas inflow rate is much weaker than for high
($e>$0.4) eccentric systems.
The number and quality of observations needed to identify the orbital periodicity
is challenging for upcoming all-sky surveys, even assuming highly modulated inflow rates
for $e\sim 0.6$ systems. This is simply because of the high number of
pointings per orbit needed (which will be the main limitation of eROSITA) and the required flux limit
of the instrument (which limits ongoing surveys like MAXI).
Detecting a PTA counterpart in a gaseous environment will thus be likely biased towards eccentric
systems that accrete the instreaming gas on time scales comparable to the binary period.
If, however, in the above scenario the time scale for the accretion of the mini--disc is long compared
to the binary period, the instreaming gas periodicity might be diluted in the process
of angular momentum redistribution within the disk, and might not be reflected
in the accretion luminosity onto the two MBHs.
In this case, stable, broad double $K \alpha$ lines arising from the inner portion of the two mini--discs
might be detectable by ATHENA \cite{Sesana11}, and could provide meaningful measurements of the
two MBH spins. If, on the contrary, the gas is instantaneously accreted, there will be no detectable
lines and the general contribution from of gap region to the emitted luminosity will likely be negligible;
only if the spread in inflowing angular momentum
is very large, colliding caustics can form small inviscid accretion discs~\cite{beloborodov2001,zalamea2009}
making the MBHs variable, luminous X--Ray sources.
Looking at circular binaries, \cite{tanaka11} have shown that gas can continue
to leak towards the MBHB even after decoupling, finding surface densities as large as
$\Sigma\sim 10^4 \, {\rm g/cm^2}$ at $\sim 40 \,GM/c^2$ for a total MBHB mass of $10^9 \msun$. While for eccentric
systems the width of the gap is generally larger than for circular ones,
their Green function approach is valid nonetheless, as long as the leaking
gas is ripped out of the edge of the circumbinary disc on dominantly radial
orbits, i.e. having a much lower angular momentum than the inner edge of the
circumbinary disc. Their results (see, e.g. their figure~4) would not change
much qualitatively, if they instead used a larger cavity to begin with and
then adapted the inner boundary condition to a function $\delta_{a}(e(t))$
rather than using $2\, a(t)$ .
However, due to gas from the decoupled disc being required
to diffuse over larger distances while the gas is also colder at the same time thus less diffusive, it is expected that the total
amount of fossil gas is smaller the higher the eccentricity.
\section{Conclusions}
\label{sec:conclusions}
We have shown that the excitation of MBHB eccentricity in both star and gas dominated
environments is a robust prediction of state of the art dynamical evolution models. In
the gas dominated case, limiting eccentricity values $e_{\rm limit} \ge 0.6$ are expected
following excitation by a cold dense circumbinary disk, whereas stellar dynamics can easily produce
eccentricities higher than 0.9. As a consequence, PTA sources, many of which are at the evolutionary
stage in which the environment still plays a role in their dynamical evolution, are expected to be
significantly eccentric. Many binaries producing residuals at a $10$ns level (in the reach of future PTA
campaigns) have $e>0.1$. The impact of such high eccentricities on the PTA signal should be
appropriately taken into account when developing dedicated data analysis algorithms.
For ELISA/NGO sources, which have long since decoupled from their environment, we have shown the
distribution of the residual eccentricity expected for both environments.
As the latter are disjoint at the high-eccentricity tail, enough GW observation statistics
could distinguish among the different scenarios, providing valuable information on
the environment of coalescing MBHBs. As far as multimessenger signals are concerned,
we did not treat electromagnetic counterparts to ELISA/NGO events and focused on PTA observations.
In this latter case, the electromagnetic signal is generally bright, due to the high mass and
relatively small luminosity distance of the sources, making certain distinct features
detectable by upcoming observatories. We have shown that counterpart identification is likely to
be bias towards eccentric MBHB in periodicity searches, whereas, depending on the angular
momentum of the instreaming material feeding the mini--discs, peculiar double $K\alpha$ lines
might be identifiable by future hard X--ray deep spectroscopy.
\ack
The authors thank Takamitsu Tanaka for helpful conversations and Massimo Dotti for carefully reading the manuscript and giving detailed comments.
\bibliographystyle{iopart-num}
|
\section{Introduction\label{intro}}
\label{sec:Intro}
The description of the hadronic excitation spectrum re\-mains a major
challenge in strong interaction theory. In spite of recent progress in
unquenched lattice QCD access to excited states is still very
limited~\cite{Edwards,Huey-Wen}. Therefore it seems worthwhile to improve
upon constituent quark model descriptions, which in view of the light quark\linebreak
masses (even taken as effective constituent masses) have to be formulated in
terms of relativistically covariant equations of motion. About a decade ago we
formulated such a quark model for baryons, see
\cite{LoeMePe1,LoeMePe2,LoeMePe3} on the basis of an instantaneous formulation
of the Bethe-Salpeter equation. In this model the quark interactions reflect a
string-like description of quark confinement through a confinement potential
rising linear\-ly with interquark distances as well as a spin-flavour
dependent interaction on the basis of instanton effects, which explains the
major spin-dependent splittings in the baryon spec\-trum.
Such a model description should offer an efficient description of masses
(resonance positions), static properties such as magnetic moments, charge
radii, electroweak amplitudes (form factors and helicity amplitudes) with only
a few model parameters. As such they also offer a framework which can be used
to judge in how far certain features could be considered to be exotic. This
concerns \emph{e.g.} phenomenological evidence for states with properties that
can not be accounted for in terms of excitations of quark degrees of freedom,
as is at the heart of any constituent quark model, but instead requires
additional degrees of freedom as \emph{e.g.} reflected by hadronic
interactions.
A satisfactory description of the major features in the light-flavoured
baryonic mass spectrum could indeed be obtained. These include
\begin{itemize}
\item
the linear Regge trajectories with an universal slope for all flavours
including states up to total angular momenta of $J=\frac{15}{2}$ and excitation
energies up to 3 GeV, see \cite{LoeMePe2,LoeMePe3};
\item
the position of the Roper-resonance and three other positive parity excited
nucleon states well below all other states of this kind. These can be largely
accounted for by the instan\-ton-induced force, the\linebreak strength of which was
chosen to reproduce the ground state $N-\Delta$ splitting
\cite{LoeMePe2,LoeMePe3};
\item
a plethora of electroweak properties which can be explained without
introducing any additional parameters, see \cite{Merten,vCauteren,Haupt}.
\end{itemize}
Nevertheless some specific discrepancies remain; most pro\-mi\-nent are:
\begin{itemize}
\item
the conspicuously low position as well as the decay properties of the
negative parity $\Lambda_{\frac{1}{2}^-}(1405)$ resonance; The calculated mass
of this state exceeds the experimental value by more than 100 MeV;
\item
there is experimental evidence~\cite{PDG} for excited negative parity
$\Delta$-resonances well below 2 GeV which can not be accounted for by the
quark model mentioned above, see fig.~\ref{Intro_Fig1},
\begin{figure}[!htb]
\centering
\includegraphics[width=\linewidth]{Deltas_SkalarConf_DiracA_Loering.eps}
\caption{
Discrepancies in the $\Delta$ mass spectrum: The left part of each
column represents the results obtained in model $\mathcal{A}$
of~\cite{LoeMePe2} in comparison with experimental data from the Particle Data
Group~\cite{PDG} (right side of each column), where lines are the resonance
position (mass) with the mass uncertainty represented by a shaded box and the
rating of~\cite{PDG} indicated by stars. $J$ and $\pi$ denote total angular
momentum and parity, respectively. Small differences with respect to the
results from fig. 3 of~\cite{LoeMePe2} are due to the fact that we obtained
increased numerical accuracy by diagonalising the resulting Salpeter
Hamiltonian in larger model spaces, see section~\ref{subsec:BSAppr} for
details.\label{Intro_Fig1}}
\end{figure}
nor by any other constituent quark model we are aware of;
\item
The mass of the positive parity $\Delta_{\frac{3}{2}^+}(1600)$ resonance,
see also fig.~\ref{Intro_Fig1}, the low value of which with respect to other
excited states of this kind can not be traced back to instan\-ton-in\-duced
effects, since these are absent for flavour symmetric states.
\end{itemize}
We therefore want to explore whether these deficiencies are inherent to the
constituent quark model itself or can be overcome by the introduction of an
additional quark interaction which improves upon the issues mentioned above
without deteriorating the excellent description of the majority of the other
states. In view of the fact that the discrepancies mainly affect the
$\Delta$-spectrum, this additional interaction is likely to be flavour
dependent. An obvious candidate in this respect would be a single pseudoscalar
meson exchange potential as has been used as a basis of an effective
spin-flavour dependent quark interaction very successfully by the Graz-group
\cite{Glozman1996,Glozman1997,Glozman1998_1
Glozman1998_2,Theussl,Glantschnig,Melde08,Plessas}.
In the present paper we shall investigate various implementations of the
coordinate (or momentum) dependence of such interactions. The paper is
organised as follows: After a brief recapitulation of the ingredients and
basic equations of our Bethe-Salpeter model (for more details
see~\cite{LoeMePe1}) in section~\ref{sec:BSmodel}, we discuss in
section~\ref{sec:BSint} the form and the parameters of the effective quark
interactions used in this paper. Section~\ref{sec:BSres} contains the results
and a discussion of the baryon mass spectra in comparison to the results
obtained before~\cite{LoeMePe2,LoeMePe3}. In
section~\ref{sec:electromagneticFF} we present some results on ground state
form factors before concluding in section~\ref{sec:Conc}.
\section{Bethe-Salpeter model}
\label{sec:BSmodel}
\subsection{Bound state Bethe-Salpeter amplitudes}
\label{subsec:BSBS}
The basic quantity describing three-quark bound states is the Bethe-Salpeter
amplitude $\chi$ defined in position space through
\begin{eqnarray}
\label{eq:BSmodel1}
\lefteqn{
\chi_{\bar{P}\,a_1 a_2 a_3}(x_1,x_2,x_3)
}
\nonumber\\
&&=
\langle 0|
T\,\Psi_{a_1}(x_1)\Psi_{a_2}(x_2)\Psi_{a_3}(x_3)
|\bar P\rangle\,,
\end{eqnarray}
where $T$ is the time ordering operator, $\left|\bar P\right\rangle$
represents the bound-state with total 4 momentum $\bar P^2=M^2$ of a baryon
with mass $M$\,, $\left|0\right\rangle$ is the physical vacuum and
$\Psi_{a_i}(x_i)$ denotes single quark-field operators with multi-indices
$a_i$ in Dirac, colour and flavour space. Because of translational invariance
below we shall exclusively use relative Jacobi coordinates $p_\xi, p_\eta$ in
momentum space. The Fourier transform of the Bethe-Salpeter amplitudes:
$\chi_{\bar P\,a_1a_2a_3}(p_{\xi},p_{\eta})$ are then determined by the
homogeneous \\
Bethe-Salpeter equation compactly written as
\begin{equation}
\label{eq:BSmodel2}
\chi_{\bar P}
=
-\textrm{i}\,
G_{0\,,P} \left(
K^{(3)}_{\bar P}
+
\bar{K}^{(2)}_{\bar P}
\right)\,
\chi_{\bar P}\,,
\end{equation}
where $K^{(3)}_P$ represents the irreducible 3-quark-kernel and where
$\bar{K}^{(2)}_P$ is defined by
\begin{eqnarray}
\label{eq:BSmodel3}
\lefteqn{
\bar{K}^{(2)}_P(p_{\xi},p_{\eta};p_{\xi}',p_{\eta}')
=
\sum_{k=1}^3
(2\pi)^4\,\delta^{(4)}(p_{\eta_k} - p_{\eta_k}')\,
}
\nonumber\\
&&
\times K^{(2)}_{(\frac{2}{3}P+p_{\eta_k})}
(p_{\xi_k},p_{\xi_k}')
\otimes
\left(S_F^3\right)^{-1}\!\!\left(\textstyle\frac{P}{3}-p_{\eta_k}\right)
\end{eqnarray}
in terms of the irreducible two-body interaction kernel
\\
$K^{(2)}_{\!(\!\frac{2}{3}P+p_{\eta_k}\!)}$ for each quark pair labeled by the
odd-particle index $k$.
Furthermore $G_{0,P}$ is the free 3-quark fermion propagator defined as
\begin{eqnarray}
\label{eq:BSmodel4}
\lefteqn{G_{0,P}(p_{\xi},p_{\eta};p_{\xi}',p_{\eta}') =}
\nonumber\\
&&
(2\pi)^8\,\delta^{(4)}(p_{\xi}-p_{\xi}')\,\delta^{(4)}(p_{\eta}-p_{\eta}')
\nonumber\\
&&
\times S^1_F(\textstyle\frac{P}{3}\!+\!p_{\xi}\!+\!\frac{p_{\eta}}{2})
\!\otimes\!
S^2_F(\frac{P}{3}\!-\!p_{\xi}\!+\!\frac{p_{\eta}}{2})
\!\otimes\!
S^3_F(\frac{P}{3}\!-\!p_{\eta})
\end{eqnarray}
in terms of full single quark propagators $S^i_F$\,.
\subsection{Model assumptions}
\label{subsec:BSAppr}
In view of the fact that the interaction kernels and the propagators are sums
of infinitely many Feynman diagrams, in order to arrive at a tractable model
we make the following assumptions, mainly with the goal to stay in close
contact with the quite successful non-relativistic constituent quark model:
\begin{itemize}
\item
The full propagators $S^i_F$ are replaced by Feynman
pro\-pagators of the free form
\begin{equation}
\label{eq:BSmodel5}
S^i_F(p)
\overset{!}{=}
\frac{\textrm{i}}{p\dslash - m_i + \textrm{i}\,\varepsilon}\,,
\end{equation}
tacitly assuming that at least some part of the self-energy can effectively
be subsumed in an effective constituent quark mass $m_i$ which then is a
parameter of the model.
\item
Obviously this does not account for confinement: This is assumed to be
implemented in the form of an instantaneous interaction kernel which in the
rest frame of the baryon is described by an unretarded potential $V^{(3)}$\,:
\begin{align}
\label{eq:BSmodel6}
K^{(3)}_P\left(p_\xi,p_\eta;{p_\xi}',{p_\eta}'\right) &
\Big|_{P=(M,\vec 0)}
\nonumber\\
\overset{!}{=} \,&
V^{(3)}\left(\vec p_\xi,\vec p_\eta; {\vec p_\xi}',{\vec p_\eta}'\right)\,.
\end{align}
Likewise we assume that two-quark interaction kernels in the
rest frame of the baryon are described by
2-body potentials $V^{(2)}$\,:
\begin{align}
\label{eq:BSmodel7}
K^{(2)}_{\frac{2}{3}P+p_{\eta_k}}\left(p_{\xi_k};{p_{\xi_k}}'\right) &
\Big|_{P=(M,\vec 0)}
\nonumber\\
\overset{!}{=} \,&
V^{(2)}\left(\vec p_{\xi_k}; {\vec p_{\xi_k}}'\right)\,.
\end{align}
\end{itemize}
Then, with a perturbative elimination of retardation effects, which arise due
to the genuine two-body interactions, see ~\cite{LoeMePe1} for details, one
can derive an equation for the (projected) Salpeter amplitude
\begin{equation}
\label{eq:BSmodel8}
\Phi^\Lambda_M(\vec p_\xi,\vec p_\eta)
=
\Lambda_+(\vec p_\xi,p_\eta)
\int
\frac{\textrm{d}p^0_\xi}{2\pi}
\frac{\textrm{d}p^0_\eta}{2\pi}
\chi_M(p_\xi,p_\eta)\,,
\end{equation}
where
\begin{eqnarray}
\label{eq:BSmodel9}
\Lambda_{\pm}(\vec p_\xi,p_\eta)
:=
\Lambda^+(\vec p_1)
\otimes
\Lambda^+(\vec p_2)
\otimes
\Lambda^+(\vec p_3)
\nonumber\\
\pm\,\,
\Lambda^-(\vec p_1)
\otimes
\Lambda^-(\vec p_2)
\otimes
\Lambda^-(\vec p_3)
\end{eqnarray}
with
\begin{equation}
\label{eq:BSmodel10}
\Lambda^{\pm}(\vec p)
=
\sum_f \Lambda^{\pm}_{m_f}(\vec p) \otimes \mathcal{P}_f\,.
\end{equation}
Here
\begin{equation}
\label{eq:BSmodel11}
\Lambda^\pm_m(\vec p) := \frac{\omega_m(\vec p) \pm H_m(\vec
p)}{2\,\omega_m(\vec p)}
\end{equation}
are projection operators on positive and negative energy states and
$\mathcal{P}_f$ projects on quark flavour $f$\,. Furthermore
the quark energy is given by $\omega(\vec p)=\sqrt{|\vec p|^2 + m^2}$ and
\begin{equation}
\label{eq:BSmodel12}
H_m(\vec p) := \gamma^0\left(\vec \gamma \cdot \vec p + m\right)
\end{equation}
is the Dirac Hamilton operator. As shown in detail in~\cite{LoeMePe1} this
equation can be written in the form of an eigenvalue problem
\begin{equation}
\label{eq:BSmodel13}
\mathcal{H}\,\Phi^\Lambda_M = M\,\Phi^\Lambda_m
\end{equation}
for the projected Salpeter amplitude, where the eigenvalues are the baryon
masses $M$\,. The Salpeter Hamilton operator is given by
\begin{eqnarray}
\label{eq:BSmodel14}
\lefteqn{
\left[\mathcal{H}\,\Phi^\Lambda_M\right](\vec p_\xi,\vec p_\eta)
=
\mathcal{H}_0(\vec p_\xi,\vec p_\eta)\,\Phi^\Lambda_M(\vec p_\xi,\vec p_\eta)
}
\nonumber\\
&+&
\Lambda_+(\vec p_\xi,\vec p_\eta)\,
\gamma^0\!\otimes\!\gamma^0\!\otimes\!\gamma^0\!\otimes\!
\int
\frac{\textrm{d}^3{p_\xi}'}{(2\pi)^3}
\frac{\textrm{d}^3{p_\eta}'}{(2\pi)^3}
\nonumber\\
&&
\hspace*{3em}
V^{(3)}(\vec p_\xi,\vec p_\eta; {\vec p_\xi}',{\vec p_\eta}')
\,\Phi^\Lambda_M({\vec p_\xi}',{\vec p_\eta}')
\nonumber\\
&+&
\Lambda_-(\vec p_\xi,\vec p_\eta)\,
\gamma^0\!\otimes\!\gamma^0\!\otimes\!\mathds{1}
\int
\frac{\textrm{d}^3{p_\xi}'}{(2\pi)^3}
\nonumber\\
&&
\hspace*{3em}
\,V^{(2)}(\vec p_\xi; {\vec p_\xi}')\otimes\!\mathds{1}
\,\Phi^\Lambda_M({\vec p_\xi}',{\vec p_\eta})
\\
&+&\textrm{corresponding quark interations (23) and (31)}\,\nonumber,
\end{eqnarray}
where $\mathcal{H}_0$ denotes the free three-quark Hamilton operator as a sum
of the corresponding single particle Dirac Hamilton operators.
The eigenvalue problem of Eq.~(\ref{eq:BSmodel13}) is solved by numerical
diagonalisation in a large but finite basis of oscillator states up to an
oscillator quantum number $N_{\textrm{max}}$\,. In previous
calculations~\cite{LoeMePe2,LoeMePe3} at least $N_{\textrm{max}}=12$ was
used. All the results in the present paper were obtained with at least
$N_{\textrm{max}}=18$\,, which, although computer time consuming, has the
advantage that for all states the independence of the numerical results on the
oscillator functions length scale in some scaling window could be warranted
and that all could be calculated with a universal value for this length
scale. This is a technical advantage when calculating electroweak amplitudes.
\section{Model Interactions}
\label{sec:BSint}
Below we specify the interaction potentials $V^{(3)}$ and $V^{(2)}$ used in
Eq.~(\ref{eq:BSmodel14})\,. These include a confinement potential, the
instanton induced two-quark interaction as has been used
before~\cite{LoeMePe2,LoeMePe3} and the new phenomenological potential
inspired by pseudoscalar meson exchange.
\subsection{Confinement}
\label{subsec:Conf}
Confinement is implemented by subjecting the quarks to a potential which rises
linearly with interquark distances, supplemented by an appropriate three
particle Dirac\linebreak structure $\Gamma$. The potential contains two parameters: the
off-set $a$ and the slope $b$ and is assumed to be of the following form in
coordinate space
\begin{equation}
\label{eq:conf}
V^{(3)}_{\textrm{conf}}(\vec x_1,\vec x_2,\vec x_3)\,
=
3\,a\,\Gamma_o
+ b \sum_{i<j} \left|\vec x_i -\vec x_j\right|\,\Gamma_s
\end{equation}
where $\Gamma_o$ and $\Gamma_s$ are suitably chosen Dirac
structures. Alternatively we can consider the linear potential to be treated
as a two-body kernel as will be used below.
\subsection{Instanton induced interaction}
Instanton effects leads to an effective quark-quark interaction, which for
quark pairs in baryons can be written in coordinate space as
\begin{eqnarray}
\label{eq:III}
\lefteqn{
\mathcal{V}^{(2)}_{\textrm{III}}(x_1,x_2;x_1'x_2')
=
V^{(2)}_{\textrm{III}}(\vec x_1 - \vec x_2)
}
\nonumber\\
&&\hspace*{2.5em}
\times\,
\delta(x_1^0-x_2^0)\,
\delta^{(4)}(x_1-x_1')\,
\delta^{(4)}(x_2-x_2')
\end{eqnarray}
with
\begin{eqnarray}
\label{eq:III2}
V^{(2)}_{\textrm{III}}(\vec x)\,
&=&
-4v(\vec x)
\left(
\mathds{1} \!\otimes\! \mathds{1} + \gamma^5 \!\otimes\! \gamma^5
\right)
\mathcal{P}^\mathcal{D}_{S_{12}=0}
\nonumber\\
&&
\hspace*{-2.5em}\!\otimes\!
\left(
g_{nn}\,\mathcal{P}^\mathcal{F}_\mathcal{A}(nn)
+
g_{ns}\,\mathcal{P}^\mathcal{F}_\mathcal{A}(ns)
\right)\,,
\end{eqnarray}
where $\mathcal{P}^\mathcal{D}_{S_{12}=0}$ is a projector on spin-singlet
states and $\mathcal{P}^{\mathcal{F}}_\mathcal{A}(f_1 f_2)$ projects
on flavour-antisymmetric quark pairs with fla\-vours $f_1$ and $f_2$.
Although the two couplings $g_{nn}$ and $g_{ns}$ are in principle determined
by integrals over instanton densities, these are treated as free parameters
here. As it stands this is a contact interaction, which for our purpose is
regularised by replacing the coordinate space dependence by a Gaussian
\begin{equation}
\label{eq:III3}
v_\lambda(\vec x)
=
\frac{1}{\lambda^3\,\pi^{\frac{3}{2}}}\,
\exp\left(-\tfrac{|\vec x|^2}{\lambda^2}\right)\,.
\end{equation}
The effective range parameter $\lambda$ is assumed to be flavour independent
and enters as an additional parameter.
\subsection{An additional flavour dependent interaction}
\label{subsec:Afdi}
The coupling of spin-$\frac{1}{2}$ fermions to a flavour nonet of
pseudoscalar meson fields is
given by an interaction Lagrange density
\begin{equation}
\label{eq:OBEps}
\mathcal{L}_I^{\textnormal{(ps)}}
=
-\mathrm{i}
\sum_{a=0}^8
g_a\,\bar{\psi}\,\gamma^5\,\lambda^a\,\psi\,\phi^a,
\end{equation}
in the case of so-called pseudoscalar coupling and by
\begin{equation}
\label{eq:OBEpv}
\mathcal{L}_I^{\textnormal{(pv)}}
=
-\sum_{a=0}^8
\frac{g_a}{2m}\,\bar{\psi}\,\gamma^5\,\gamma^{\mu}\,
\lambda^a\,\psi\,\partial_{\mu}\phi^a.
\end{equation}
for pseudovector coupling. Here $\psi$ represents the quark fields
with mass $m$ and $\phi^a$ the pseudoscalar meson fields with mass $\mu_a$
where the flavour index $a=\pi^{\pm\,,0},\eta^0_8,$ $\eta_1^0$ $K^\pm,$ $K^0,$
$\bar K^0$\,. The flavour dependence is represented by the usu\-al Gell-Mann
matrices $\lambda^a, a=1,\ldots,8$\,; $\lambda^0$ is proportional to the
identity operator in flavour space normalised to
$\textrm{Tr}((\lambda^0)^2)=2$.
A standard application of the Feynman rules, see \emph{e.g.} ~\cite{CaDuElHaSiSp}
then leads to the second order scattering-matrix element
$\mathcal{M}^{(2)}$ given in the CM-system by the expressions
\begin{eqnarray}
\label{eq:OBEM2ps}
\lefteqn{
\mathrm{i}
\mathcal{M}_{\textnormal{(ps)}}^{(2)}(k_0,\vec{k})
}
\nonumber\\
&=&
\sum_{a,b} g_a^2
\bigg[
\bar{\psi}(\vec{p}')(-i\gamma^5)\lambda^a\,\psi(\vec{p})\,
D^{ab}(k_0,\vec{k})
\nonumber\\
&&
\hspace*{1em}\times\bar{\psi}(-\vec{p}')
(-\mathrm{i}\gamma^5)
\lambda^b
\psi(-\vec{p})
\bigg]
\nonumber\\
&=:&
-
\sum_{a,b} g_a^2
D^{ab}(k_0,\vec{k})
\left[\lambda^a\,\gamma^5\right]
\! \otimes\!
\left[\lambda^b\,\gamma^5\right]
\nonumber\\
&=:&
\mathrm{i}
[\bar{\psi}(-\vec{p}')\!\otimes\!\bar{\psi}(\vec{p}')]
V_{(ps)}(k_0,\vec{k})
[\psi(-\vec{p})\!\otimes\!\psi(\vec{p})]
\end{eqnarray}
and
\begin{eqnarray}
\label{eq:OBEM2pv}
\lefteqn{
\mathrm{i}\,\mathcal{M}_{\textnormal{(pv)}}^{(2)}(k_0,\vec{k})
}
\nonumber\\
&=&
\sum_{a,b} \frac{g_a^2}{4m^2}
\bigg[
\bar{\psi}(\vec{p}')\gamma^5\gamma^{\mu}(-ik_{\mu})
\lambda_a\psi(\vec{p})\,
D^{ab}(k_0,\vec{k})
\nonumber\\
&&
\hspace*{1em}\times
\bar{\psi}(-\vec{p}')\gamma^5\gamma^{\nu}(-i(-k_{\nu}))
\lambda_b\psi(-\vec{p})
\bigg]
\nonumber\\
&=:&
\sum_{a,b}
\frac{g_a^2}{4m^2}\,
D^{ab}(k_0,\vec{k})\,k_{\mu}k_{\nu}\!
\left[\lambda^a\gamma^5\gamma^{\mu}\right]\!\!\otimes\!\!
\left[\lambda^b\gamma^5\gamma^{\nu}\right]
\nonumber\\
&:=&
\mathrm{i}\,
[\bar{\psi}(-\vec{p}')\!\otimes\!\bar{\psi}(\vec{p}')]
V_{(pv)}(k_0,\vec{k})
[\psi(-\vec{p})\!\otimes\!\psi(\vec{p})]
\end{eqnarray}
in case of pseudoscalar and pseudovector coupling, respectively. Here the
meson propagator is given by
\begin{equation}
\label{eq:prop}
D^{ab}(k_0,\vec{k})
=
\frac{\mathrm{i}\delta_{ab}}{k_0^2-|\vec{k}|^2-\mu_a^2}\,,
\end{equation}
with $k_{\mu}:=p_{\mu}'-p_{\mu}$ the momentum transfer.
In instantaneous approximation we set $k_0 = 0$. From
Eqs. [\ref{eq:OBEM2ps},{\ref{eq:OBEM2pv}] we extract the corresponding
potentials in momentum space
\begin{eqnarray}
\label{eq:OBEVps}
V_{(ps)}^{(2)}(\vec{k})
=
\sum_a g_a^2
[\lambda^a\!\otimes\!\lambda^a]
\frac{1}{|\vec{k}|^2+\mu_a^2}
\bigg[\gamma^5\!\otimes\!\gamma^5\bigg]
\end{eqnarray}
for pseudoscalar coupling and
\begin{eqnarray}
\label{eq:OBEVpv}
\lefteqn{
V_{(pv)}^{(2)}(\vec{k})
}
\\
&=&
\sum_a \frac{g_a^2}{4m^2}
[\lambda^a\!\otimes\!\lambda^a]
\frac{-1}{|\vec{k}|^2+\mu_a^2}
\Big[\!
\big(\gamma^5\vec{\gamma}\cdot\vec{k}\big)
\!\otimes\!
\big(\gamma^5\boldsymbol{\gamma}\cdot\vec{k}\big)
\!\Big]
\nonumber\\
&=&
\sum_a
\frac{g_a^2}{4m^2}
[\lambda^a\!\otimes\!\lambda^a]
\frac{-|\vec{k}|^2}{|\vec{k}|^2+\mu_a^2}
\Big[\!
\big(\gamma^5\vec{\gamma}\cdot\widehat{\vec{k}}\big)
\!\otimes\!
\big(\gamma^5\boldsymbol{\gamma}\cdot\widehat{\vec{k}}\big)
\!\Big]
\nonumber
\end{eqnarray}
for pseudovector coupling, where $\widehat{\vec k} := \frac{\vec k}{|\vec
k|}$\,.
As it stands, the expression for the potential in the instantaneous
approximation for pseudoscalar coupling\newline leads, after Fourier transformation,
to a local Yukawa potential in configuration space with the usual range given
by the mass of the exchanged pseudoscalar meson. For pseudovector coupling the
non-relativistic approximation to the Fourier transform leads to the usual
spin-spin contact interaction together with the usual tensor force. In the
simplest form adopted by the Graz group
\cite{Glozman1996,Glozman1997,Glozman1998_1
Theussl,Glantschnig,Melde08,Plessas}
the latter were ignored, in addition the contact term was regularised by a
Gaussian function and the Yukawa terms were regularised to avoid singularities
at the origin.
In view of this and the instantaneous approximation we decided to parametrise
the new flavour dependent interaction purely phenomenologically as a local
potential in configuration space, its simple form given by
\begin{equation}
\label{eq:OBEVreg}
V^{(2)}(\vec x)
=
\sum_a g_a^2\,
\left[
\lambda^a\,\gamma^5 \!\otimes\! \lambda^a\,\gamma^5
\right]\,
v_{\lambda_a}(\vec x)
\end{equation}
where $v_\lambda(\vec x)$ is the Gaussian form given in Eq.~\ref{eq:III3}\,.
Other Dirac structures, such as
$\left[\gamma^5 \vec \gamma\cdot \widehat{\vec x}\,
\!\otimes\!
\gamma^5 \vec \gamma\cdot \widehat{\vec x}\right]$
were tried, but were found to be less effective. Results for the meson
exchange form of the interaction as given by
Eqs.~(\ref{eq:OBEVps},\ref{eq:OBEVpv}) will be briefly discussed in
section~\ref{sec:Conc}.
\begin{table}[!htb]
\centering
\caption{
The list of baryon resonances of which the masses were used to determine the
model parameters in a least-squares fit where every resonance was attributed a
weight reciprocal to its uncertainty in its position as given in \cite{PDG}.
Nominal masses are given in MeV.
\label{tab:Tab_3}
}
\begin{tabular}{
@{\hspace{0pt}}l@{\hspace{2pt}}l@{\hspace{2pt}}l@{\hspace{2pt}
l@{\hspace{6pt}}l@{\hspace{0pt}
}
\toprule
$\Delta$ & $N$ & $\Lambda / \Sigma$ & $\Xi$ & $\Omega$ \\
\midrule
$S_{31}(1620)$ & $S_{11}(1535)$ & $S_{21}(1620)$ & $S_{11}(1309)$ & \\
$S_{31}(1900)$ & $S_{11}(1650)$ & $S_{21}(1750)$ & & \\
\midrule
$P_{31}(1750)$ & $P_{11}(939)$ & $P_{01}(1116)$ & $P_{13}(1530)$ & $P_{01}(1672)$\\
$P_{31}(1910)$ & $P_{11}(1440)$ & $P_{21}(1289)$ & & \\
$P_{33}(1232)$ & & $P_{01}(1600)$ & & \\
$P_{33}(1600)$ & & $P_{21}(1660)$ & & \\
$P_{33}(1920)$ & & & & \\
\midrule
$D_{33}(1700)$ & & & $D_{13}(1820)$ & \\
$D_{33}(1940)$ & & & & \\
$D_{35}(1930)$ & & & & \\
\midrule
$F_{35}(1905)$ & & & & \\
$F_{35}(2000)$ & & & & \\
$F_{37}(1950)$ & & & & \\
\midrule
$G_{37}(2200)$ & & & & \\
$G_{39}(2400)$ & & & & \\
\midrule
$H_{39}(2300)$ & & & & \\
$H_{3\,11}(2420)\!\!\!$ & & & & \\
\midrule
$K_{3\,15}(2950)\!\!\!$ & & & & \\
\bottomrule
\end{tabular}
\end{table}
\section{Mass spectra}
\label{sec:BSres}
\subsection{Model parameters}
\label{subsec:parameters}
The resulting baryon mass spectra were obtained by fitting the parameters of
the model, \emph{viz.} the offset $a$ and slope $b$ of the confinement
potential, the constituent quark masses $m_n=m_u=m_d$ and $m_s$, the strengths
of the instanton induced force, $g_{nn}$ and $g_{ns}$ as well as the strengths
of the additional flavour dependent interaction, given by $g_8$ and $g_0$ for
flavour octet and flavour singlet exchange (thus assuming $SU(3)$ symmetry) to
a selection of baryon resonances, see table~\ref{tab:Tab_3}\,.
The range $\lambda$ given to the instanton induced force was kept
to the value used in~\cite{LoeMePe2,LoeMePe3} and is roughly in accordance
with typical instanton sizes. The optimal value for the range of the
additional flavour dependent interaction was found to be $\lambda_8 = \lambda_0
\approx 0.25\,\textnormal{fm}$ and thus turned out to be of rather short range.
A comparison of the parameters obtained with the parameters of model
$\mathcal{A}$ of~\cite{LoeMePe2,LoeMePe3} is given in table~\ref{tab:par}\,.
\begin{table}[!htb]
\centering
\caption{Model parameters for the current model $\mathcal{C}$ in
comparison to those of model $\mathcal{A}$ of~\cite{LoeMePe2,LoeMePe3}.
Some of the parameters have been slightly changed with respect to
the original values (listed in brackets) of~\cite{LoeMePe2,LoeMePe3},
since the calculation has been performed with higher numerical accuracy
by taking more basis states in the diagonalisation of the Salpeter
Hamiltonian, see also text.
\label{tab:par}}
\begin{tabular}{l@{\hspace{10pt}}l@{\hspace{10pt}}r@{\hspace{10pt}}r}
\toprule
parameter & & model $\mathcal{C}$ & model $\mathcal{A}$ \\
\midrule
masses & $m_n$ [MeV] & 325.0 & 330.0 \\
& $m_s$ [MeV] & 600.0 & 670.0 \\
\midrule
\multirow{2}{*}{confinement} & \multirow{2}{*}{$a$ [MeV]} & \multirow{2}{*}{-366.78} & -734.6 \\
& & & [-744.0] \\
& \multirow{2}{*}{$b$ [MeV/fm]} & \multirow{2}{*}{212.81} & 453.6\\
& & & [440.0] \\
\midrule
\multirow{2}{*}{instanton} & \multirow{2}{*}{$g_{nn}$ [MeV fm$^3$]} & \multirow{2}{*}{341.49} & 130.3 \\
& & & [136.0] \\
\multirow{2}{*}{induced} & \multirow{2}{*}{$g_{ns}$ [MeV fm$^3$]} & \multirow{2}{*}{273.55} & 81.8 \\
& & & [96.0] \\
interaction & $\lambda$ [fm] & 0.4 & 0.4 \\
\midrule
octet & $\frac{g_{8}^2}{4\pi}$ [MeV fm$^3$] & 100.86 & -- \\
exchange & & & \\
singlet & $\frac{g_{0}^2}{4\pi}$ [MeV fm$^3$] & 1897.43 & -- \\
exchange & & & \\
& $\lambda_{8}=\lambda_0$ [fm] & 0.25 & -- \\
\bottomrule
\end{tabular}
\end{table}
\begin{table}[!htb]
\centering
\caption{
Multiplet decomposition of amplitudes of negative parity
$\Delta$-resonances. For each amplitude the contribution to the Salpeter norm,
see~\cite{LoeMePe1} is given in $\%$\,, in each row the upper line and the
lower line give the positive and negative energy contribution,
respectively. States are labeled by the calculated mass and $J^\pi$ denotes
total angular momentum and parity, $^{2S+1}\mathcal{F}_{J}[D]$ label
amplitudes with spin $S$, flavour representation with dimension $\mathcal{F}$,
$SU(6)$ representation with dimension $D$\,. The dominant contribution is
underlined. \label{tab:LSD}
}
\begin{tabular}[c]{cc@{\hspace{30pt}}c@{\hspace{30pt}}rr}
\toprule
$J^\pi$ & Mass & pos. & $^{4}10_{J}[56]$ & $^{2}10_{J}[70]$ \\
& & neg. & $^{4}10_{J}[56]$ & $^{2}10_{J}[70]$ \\
\midrule
$\frac{1}{2}^-$ & 1635 & 98.9 & 6.6 & \underline{92.3} \\
& & 1.1 & 0.8 & 0.4 \\
\midrule
$\frac{1}{2}^-$ & 1932 & 98.5 & 14.2 & \underline{84.3} \\
& & 1.5 & 1.0 & 0.5 \\
\midrule
$\frac{1}{2}^-$ & 2041 & 99.0 & \underline{81.6} & 17.3 \\
& & 1.0 & 0.5 & 0.5 \\
\midrule
$\frac{3}{2}^-$ & 1592 & 98.2 & 9.9 & \underline{88.4} \\
& & 1.8 & 0.9 & 0.9 \\
\midrule
$\frac{3}{2}^-$ & 1871 & 97.9 & 24.5 & \underline{73.3} \\
& & 2.1 & 1.4 & 0.7 \\
\midrule
$\frac{3}{2}^-$ & 1944 & 98.6 & \underline{65.1} & 33.5 \\
& & 1.4 & 0.9 & 0.6 \\
\midrule
$\frac{5}{2}^-$ & 2013 & 98.8 & \underline{89.3} & 9.6 \\
& & 1.2 & 0.6 & 0.6 \\
\midrule
$\frac{5}{2}^-$ & 2136 & 99.2 & 16.8 & \underline{82.4} \\
& & 0.8 & 0.3 & 0.5 \\
\midrule
$\frac{5}{2}^-$ & 2166 & 99.1 & \underline{86.6} & 12.4 \\
& & 1.0 & 0.4 & 0.6 \\
\bottomrule
\end{tabular}
\end{table}
The parameters need some comments: In the original
paper~\cite{LoeMePe2,LoeMePe3} the Dirac structures (\emph{i.e.} the spin
dependence) of the confinement potential were taken to be
$\Gamma_0 = \frac{1}{4}(
\mathds{1}\!\otimes\!\mathds{1}\!\otimes\!\mathds{1}
+\gamma^0\!\otimes\!\gamma^0\!\otimes\!\mathds{1}
+\textrm{cycl. perm.})$
and
$
\Gamma_s = \frac{1}{2}(
-\mathds{1}\!\otimes\!\mathds{1}\!\otimes\!\mathds{1}
+\gamma^0\!\otimes\!\gamma^0\otimes\mathds{1}
+\textrm{cycl. perm.})$
for the offset and slope, respectively, and were considered to build a 3-body
kernel. In the present model including the additional octet and singlet
flavour exchange potential into account we obtained the best results with
$\Gamma_0 =
\mathds{1}\!\otimes\!\mathds{1}\!\otimes\!\mathds{1}
$
and
$
\Gamma_s =
\gamma^0\!\otimes\!\gamma^0
$
and treating the interaction corresponding to the latter term as a 2-body
interaction. This of course impedes a direct comparison of the corresponding
parameters. Furthermore it was found that the strengths of the instanton
induced interaction is roughly tripled when compared to the original
values. Note that the additional flavour exchange interaction has the same
spin-flavour dependence as parts of the former interaction. The flavour
singlet exchange could effectively be considered as an other spin dependent
part of the confinement potential. Possibly this explains the extraordinary
large coupling in this case. In summary it thus must be conceded that the
present treatment is phenomenological altogether and that here unfortunately
the relation to more fundamental QCD parameters, such as instanton couplings
and string tension is lost. Nevertheless with only 10 parameters we consider
the present treatment to be effective especially in view of its merits in the
improved description of some resonances to be discussed below.
\subsection{The $\Delta$ and the $\Omega$ spectrum}
\label{subsec:DeltaSpectrum}
In fig.~\ref{Fig_GaussDeltaPSMeson} we compare the results from the present
calculation (model $\mathcal{C}$) (right side of each column) with
experimental data from the Particle Data Group~\cite{PDG} (central in each
column) and with the results from model $\mathcal{A}$ of ~\cite{LoeMePe2}
(left side in each column)\,. The parameters used are listed in
table~\ref{tab:par}\,.
\begin{figure*}[!htb]
\centering
\includegraphics[width=\linewidth]{Gauss_DeltasPS_coupled.eps}
\caption{
Comparison of the $\Delta$-Spectrum calculated within the present model
$\mathcal{C}$ (right side of each column) with experimental data from the
Particle Data Group~\cite{PDG} (central in each column) and with the results
from model $\mathcal{A}$ of ~\cite{LoeMePe2} (left side in each column), note
the caption to fig.~\ref{Intro_Fig1}; Lines indicate the resonance position
(mass) with the mass uncertainty represented by a shaded box and the rating
of~\cite{PDG} indicated by stars. The small numbers give the mass in MeV. $J$
and $\pi$ denote total angular momentum and parity, respectively.
\label{Fig_GaussDeltaPSMeson}}
\end{figure*}
\begin{table}[!htb]
\centering
\caption{
Comparison of experimental~\cite{PDG} and calculated masses in MeV of
$\Delta$-resonances. The corresponding spectra
are shown in fig.~\ref{Fig_GaussDeltaPSMeson}\,.
\label{tab:Delta}
}
\begin{tabular}
*{6}{@{\hspace{2pt}}r
}
\toprule
\multicolumn{1}{c}{exp.} & rating & model $\mathcal{C}$ &
\multicolumn{1}{c}{exp.} & rating & model $\mathcal{C}$ \\
\midrule
$S_{31}(1620)$ & \tiny{****} & 1634 &
$S_{31}(1900)$ & \tiny{***} & 1932
\\
$S_{31}(2150)$ & \tiny{*} & 2040/2109
& &
\\
\midrule
$P_{31}(1750)$ & \tiny{*} & 1653 &
$P_{31}(1910)$ & \tiny{****} & 1888
\\
$P_{33}(1232)$ & \tiny{****} & 1231 &
$P_{33}(1600)$ & \tiny{***} & 1559
\\
$P_{33}(1920)$ & \tiny{***} & 1894/1925
& &
\\
\midrule
$D_{33}(1700)$ & \tiny{****} & 1591 &
$D_{33}(1940)$ & \tiny{*} & 1871/1944
\\
$D_{35}(1930)$ & \tiny{***} & 2012
& &
\\
\midrule
$F_{35}(1905)$ & \tiny{****} & 1897 &
$F_{35}(2000)$ & \tiny{*} & 1961
\\
$F_{37}(1950)$ & \tiny{****} & 1936 &
$F_{37}(2390)$ & \tiny{*} & many res.
\\
\midrule
$G_{37}(2200)$ & \tiny{*} & 2126/2198 &
$G_{39}(2400)$ & \tiny{**} & 2232\\
\midrule
$H_{39}(2300)$ & \tiny{**} & 2319 &
$H_{3,11}(2420)$ & \tiny{****} & 2372
\\
\midrule
$I_{3,13}(2750)$ & \tiny{**} & 2584 &
& &
\\
\midrule
$K_{3,15}(2950)$ & \tiny{**} & 2724/2909 &
& &
\\
\bottomrule
\end{tabular}
\end{table}
\begin{table}[!htb]
\centering
\caption
Comparison of experimental~\cite{PDG} and calculated masses in MeV of
$N$-resonances. The corresponding spectra
are shown in fig.~\ref{Fig_GaussNucleonPSMeson}\,.
\label{tab:Tabnine}
}
\begin{tabular}
*{6}{@{\hspace{2pt}}r}
}
\toprule
\multicolumn{1}{c}{exp.} & rating & model $\mathcal{C}$ &
\multicolumn{1}{c}{exp.} & rating & model $\mathcal{C}$ \\
\midrule
$S_{11}(1535)$ & \tiny{****} & 1484 &
$S_{11}(1650)$ & \tiny{****} & 1672
\\
$S_{11}(2090)$ & \tiny{*} & many res.&
& &
\\
\midrule
$P_{11}(939)$ & \tiny{****} & 945 &
$P_{11}(1440)$ & \tiny{****} & 1440
\\
$P_{11}(1710)$ & \tiny{***} & 1709 &
$P_{11}(2100)$ & \tiny{*} & many res.
\\
$P_{13}(1720)$ & \tiny{****} & 1703 &
$P_{13}(1900)$ & \tiny{**} & 1825\\
\midrule
$D_{13}(1520)$ & \tiny{****} & 1534 &
$D_{13}(1700)$ & \tiny{***} & 1685
\\
$D_{13}(2080)$ & \tiny{**} & many res.&
$D_{15}(1675)$ & \tiny{****} & 1667
\\
$D_{15}(2200)$ & \tiny{**} & many res.
& &
\\
\midrule
$F_{15}(1680)$ & \tiny{****} & 1761 &
$F_{15}(2000)$ & \tiny{**} & 1930/1983
\\
$F_{17}(1990)$ & \tiny{**} & 2001 &
& &
\\
\midrule
$G_{17}(2190)$ & \tiny{****} & 2011 &
$G_{19}(2250)$ & \tiny{****} & 2165
\\
\midrule
$H_{19}(2220)$ & \tiny{****} & 2192 &
& &
\\
\midrule
$I_{1,11}(2600)$ & \tiny{***} & 2377 &
& &
\\
\midrule
$K_{1,13}(2700)$ & \tiny{**} & 2543 &
& &
\\
\bottomrule
\end{tabular}
\end{table}
The spectrum of the $\Delta$ (see fig.~\ref{Fig_GaussDeltaPSMeson}) and
$\Omega$ (see right panel of fig. \ref{Fig_GaussXiOmegaPSMeson}}) resonances
is determined by the confinement potential and the flavour exchange
interaction only, since the instanton induced interaction does not act on
flavour symmetric states. Concerning the positive parity resonances we see
that in the present calculation we can now indeed account for the low position
of the $\Delta_{\frac{3}{2}^+}(1600)$ resonance. In addition the next
excitations in this channel now lie closer to 2000 MeV in better agreement
with experimental data, as is also the case for the splitting of the two
$\Delta_{\frac{1}{2}^+}$ resonances. Note, however, that these states were
included in the parameter fit. Additionally there is support for a parity
doublet $\Delta_{\frac{3}{2}^+}(1920)$ and $\Delta_{\frac{3}{2}^-}(1940)$ as
argued by~\cite{Horn}.
Likewise, we can now account for the excited negative parity resonances:
$\Delta_{\frac{1}{2}^-}(1900)$\,,$\Delta_{\frac{3}{2}^-}(1940)$ and
$\Delta_{\frac{5}{2}^-}$ $(1930)$ and even find two states in the
$\Delta_{\frac{3}{2}^-}$ channel which could correspond to the poorly
established $\Delta_{\frac{3}{2}^-}(1940)$ state.
In view of the near degeneracy of the $\Delta_{\frac{1}{2}^-}$ $(1900)$\,,
$\Delta_{\frac{3}{2}^-}$ $(1940)$ and $\Delta_{\frac{5}{2}^-}(1930)$ states it
is tempting to classify these in a non-relativistic scheme as a total spin
$S=\frac{3}{2}$, total quark angular momentum $L=1$ multiplet, which, because
of total isospin $I=\frac{3}{2}$ must then belong to a $(56,1^-)$ multiplet,
which is lowered with respect to the bulk of the other negative parity states
that in an oscillator classification would be attributed to the $N=3$ band.
Obviously this is not supported by the calculations: As table~\ref{tab:LSD}
shows, although the lowest $J^\pi =\frac{5}{2}^-$ resonance has a dominant
component in this multiplet, the second excited $J^\pi = \frac{1}{2}^-,
\frac{3}{2}^-$ resonances have dominant components in the $(70,1^-)$
multiplet; indeed the third excited states in these channels can be attributed
to the $(56,1^-)$ multiplet. Otherwise the description and in particular the
$\Delta$-Regge-trajectory are of a similar quality as in the original
model $\mathcal{A}$\,.
Concerning the $\Omega$ spectrum, see fig.~\ref{Fig_GaussXiOmegaPSMeson}
(right panel), apart from the appearance of an excited
$\Omega_{\frac{3}{2}^+}$ state at 2021 MeV no spectacular changes in the
predictions with respect to the original model $\mathcal{A}$ were found. Note
that the present model predicts that this state is almost degenerate with the
first negative parity states $\Omega_{\frac{1}{2}^-}$ at 2008 MeV and
$\Omega_{\frac{3}{2}^-}$ at 1983 MeV.
In table~\ref{tab:Delta} we have summarised the calculation of
$\Delta$-res\-on\-an\-ces.
\subsection{The $N$ spectrum}
\label{subsec:NucleonSpectrum}
In fig.~\ref{Fig_GaussNucleonPSMeson} we present the results for the nucleon
spectrum.
\begin{figure*}[!htb]
\centering
\includegraphics[width=\linewidth]{Gauss_NucleonsPS_coupled.eps}
\caption{
Comparison of the $N$-Spectrum calculated within the present model
$\mathcal{C}$ (right side of each column) with experimental data from the
Particle Data Group~\cite{PDG} (central in each column) and with the results
from model $\mathcal{A}$ of ~\cite{LoeMePe2} (left side in each column).
See also caption to fig.~\ref{Fig_GaussDeltaPSMeson}.
\label{Fig_GaussNucleonPSMeson}
}
\end{figure*}
As was the case for the $\Delta$ spectrum, comparing to the results from the
former model $\mathcal{A}$ we indeed obtain an improved description of the
position of the first excited state with the same quantum numbers as the
ground state, the so called Roper resonance, while at the same time improving
also on the position of the first excited negative parity resonances $J^\pi =
\big(\tfrac{1}{2}^-\big)_{1}$,$\big(\tfrac{1}{2}^-\big)_{2}$,
$\big(\tfrac{3}{2}^-\big)_{1}$, $\big(\tfrac{3}{2}^-\big)_{2}$ and
$\big(\tfrac{5}{2}^-\big)_{1}$\,. With the exception of the
$J^\pi=\frac{5}{2}^+$ state, which compared to model $\mathcal{A}$ is shifted
upwards by approximately 50 MeV, the description of all known excited states
is of a similar quality as that of model $\mathcal{A}$. In particular the
position of the lowest $J^\pi = \frac{7}{2}^-$ resonance is still
underestimated by more than 100 MeV.
In the following we compare the predictions obtained in model $\mathcal{C}$
for nucleon resonances with $J \le \frac{5}{2}$ and masses larger than 1.8 GeV
with new results obtained in the Bonn-Gatchina analyses as reported in
\cite{Anisovich_1,Anisovich_2}: In particular in \cite{Anisovich_1} a fourth
$J^\pi = \frac{1}{2}^+$ state was found, called $N_{\frac{1}{2}^+}(1875)$
which could correspond to our calculated state at 1893 MeV. Furthermore the
analysis contains two\linebreak $J^\pi = \frac{3}{2}^+$ states, called $N(1900)P_{13}$
and $N_{\frac{3}{2}^+}(1975)$, which might be identified with the model
$\mathcal{C}$ states calculated at 1825 MeV and 1945 MeV (or 1966 MeV),
respectively. Concerning the negative parity states, in \cite{Anisovich_2} a
new $J^\pi = \frac{1}{2}^-$ state was found ($N_{\frac{1}{2}^-}(1895)$) which
could be identified with the calculated state at 1851 MeV (or with that at
1881 MeV). In addition two $J^\pi = \frac{3}{2}^-$ states were found:
$N_{\frac{3}{2}^-}(1875)$ could correspond to the calculated states at 1853
MeV (or at 1934 MeV) and $N_{\frac{3}{2}^-}(2150)$ with one of the three
states with calculated masses 2073 MeV, 2091 MeV and 2137 MeV. The new
$N_{\frac{5}{2}^-}(2060)$ state reported in \cite{Anisovich_2} is closest to
the states calculated at 1945 MeV and at 2007 MeV. Finally, for $J^\pi =
\frac{5}{2}^+$ the analysis is ambiguous: Although a solution with a single
pole around 2.1 GeV is not excluded, solutions with 2 poles, either an
ill-defined pole in the 1800-1950 MeV mass region and one at nearly 2.2 GeV or
two close poles at approximately 2.0 GeV were found and could correspond to
the model $\mathcal{C}$ states calculated at 1930 MeV and 1983 MeV. Note that
these new resonances were not included in the parameter fit (see table~\ref{tab:Tab_3}).
An overview of the identification of nucleon resonances is given in
table~\ref{tab:Tabnine}.
\subsection{Hyperon spectra}
\label{subsec:YSpectra}
\begin{figure*}[!htb]
\centering
\includegraphics[width=\linewidth]{Gauss_LambdasPS_coupled.eps}
\caption{
Comparison of the $\Lambda$-Spectrum calculated within the present model
$\mathcal{C}$ (right side of each column) with experimental data from the
Particle Data Group~\cite{PDG} (central in each column) and with the results
from model $\mathcal{A}$ of ~\cite{LoeMePe2} (left side in each column), see
also caption to fig.~\ref{Fig_GaussDeltaPSMeson}.
\label{Fig_GaussLambdaPSMeson}
}
\end{figure*}
\begin{figure*}[!htb]
\centering
\includegraphics[width=\linewidth]{Gauss_SigmasPS_coupled.eps}
\caption{
Comparison of the $\Sigma$-Spectrum calculated within the present model
$\mathcal{C}$ (right side of each column) with experimental data from the
Particle Data Group~\cite{PDG} (central in each column) and with the results
from model $\mathcal{A}$ of ~\cite{LoeMePe2} (left side in each column), see
also caption to fig.~\ref{Fig_GaussDeltaPSMeson}.
\label{Fig_GaussSigmaPSMeson}
}
\end{figure*}
\begin{figure*}[!htb]
\centering
\psfrag{Xi}[c][c]{\Huge$\Xi$}
\psfrag{Om}[c][c]{\Huge$\Omega$}
\includegraphics[width=\linewidth]{Gauss_XisOmegaPS_coupled.eps}
\caption{
Comparison of the $\Xi$-Spectrum (first eight columns) and the
$\Omega$-Spectrum (rightmost eight columns) calculated within the present model
$\mathcal{C}$ (right side of each column) with the experimental data from the
Particle Data Group~\cite{PDG} (central in each column) and with the results
from model $\mathcal{A}$ of ~\cite{LoeMePe2} (left side in each column), see
also caption to fig.~\ref{Fig_GaussDeltaPSMeson}.
\label{Fig_GaussXiOmegaPSMeson}
}
\end{figure*}
The resulting spectra for hyperon resonances, \emph{viz.} the
$\Lambda$\,,$\Sigma$ and $\Xi$ states are depicted in
figs.~\ref{Fig_GaussLambdaPSMeson}, \ref{Fig_GaussSigmaPSMeson} and
\ref{Fig_GaussXiOmegaPSMeson}\,, respectively.
Again we indeed find an improved description of the ``Roper-like'' resonances
$\Lambda_{\frac{1}{2}^+}(1600)$ and $\Sigma_{\frac{3}{2}^+}(1660)$. Note,
however, that both were used to determine the model parameters. Concerning the
negative parity resonances, although we do find an acceptable description of
the $\Lambda$ resonances with $J^\pi = \frac{3}{2}^-\,, \frac{5}{2}^-$ and
$\frac{7}{2}^-\,$, also the new cal\-culat\-ion can not account for the low
position of the $\Lambda_{\frac{1}{2}^-}(1405)$ resonance, which now is 200
MeV below the calculated position. In our opinion this underlines the
conclusion, that this state cannot indeed be accounted for in terms of a
$q^3$ excitation alone and that its position is determined by a strong
coupling of a ``bare'' $q^3$ state to meson-baryon decay channels due to the
proximity of the $\bar{K}N$-threshold, see also
ref.~\cite{Jido,Hyodo2008_1,Hyodo2008_2} for a description of this state in a
chiral unitary approach.
Concerning the $\Xi$ resonances, with respect to model $\mathcal{A}$
of~\cite{LoeMePe3} mainly the prediction for the excited state
$\Xi(\frac{1}{2}^+)$ at 1765 MeV is 100 MeV lower. To a lesser extend this also
holds for the excited $\Xi(\frac{3}{2}^+)$ which is now predicted at 1889 MeV.
\section{Electromagnetic properties}
\label{sec:electromagneticFF}
As has been elaborated in~\cite{Merten} in lowest order the transition current
matrix element for an initial baryon state with four-momentum
$\bar{P}'=M=(M,\vec 0)$
in its rest frame and a final baryon state with four-momentum
$\bar{P}$
is given by the expression
\begin{eqnarray}
\label{eq:em1}
&&\langle \bar P | j^\mu(0) | M \rangle
=
\!-3\!\!\int\!\!\frac{\textrm{d}^4p_\xi}{(2\pi)^4}\!\!
\int\!\!\frac{\textrm{d}^4p_\eta}{(2\pi)^4}\!
\textstyle
{\bar{\Gamma}}^\Lambda_{\bar{P}}\!\left(p_\xi,p_\eta\!-\!\frac{2}{3}q\right)
\nonumber
\\
&&\hspace*{1em}
\textstyle
S^1_F\!\left(\frac{1}{3}M+\!p_\xi\!+\!\frac{1}{2}p_\eta\right)\!
\otimes\!
S^2_F\!\left(\frac{1}{3}M-\!p_\xi\!+\!\frac{1}{2}p_\eta\right)
\nonumber
\\
&&\hspace*{1em}
\textstyle
\otimes
S^3_F\!\left(\frac{1}{3}M\!-\!p_\xi\!-\!p_\eta\!+\!q\right)
\widehat q\gamma^\mu
S^3_F\!\left(\frac{1}{3}M\!-\!p_\xi\!-\!p_\eta\right)
\nonumber
\\
&&
\hspace*{10em}
\textstyle
\Gamma^\Lambda_M\left(\vec p_\xi,\vec p_\eta\right)\,.
\end{eqnarray}
Here $q$ denotes the momentum transfer, $\widehat q$ is the charge operator
and
\begin{eqnarray}
\label{eq:em2}
\lefteqn{
\Gamma^\Lambda_M\left(\vec p_\xi,\vec p_\eta\right)
:=
-\textrm{i}
\int\frac{\textrm{d}p'_\xi}{(2\pi)^4}
\int\frac{\textrm{d}p'_\eta}{(2\pi)^4}
}\nonumber
\\
&&
\hspace*{1em}\left[
V^{(3)}_\Lambda\left(\vec p_\xi,\vec p_\eta;\vec p'_\xi,\vec p'_\eta\right)
+
V^{\textrm{eff}}_\Lambda\left(\vec p_\xi,\vec p_\eta;\vec p'_\xi,\vec
p'_\eta\right)
\right]\nonumber
\\
&&
\hspace*{10em}\Phi^\lambda_M\left(\vec p'_\xi,\vec p'_\eta\right)
\end{eqnarray}
is the vertex function in the rest frame of the baryon with $V_\Lambda$ the
projection of the instantaneous interaction kernels onto the subspace of
purely positive and purely negative energy components only, see in particular
Appendix A of~\cite{Merten} for details. The vertex for a general
four-momentum on the mass shell can be obtained by an appropriate
Lorentz-boost.
Accordingly, the Sachs form factors are given by
\begin{subequations}
\begin{eqnarray}
\label{eq:em3a}
G^N_E(Q^2)
&=&
\frac{\langle N, \bar{P}, \frac{1}{2}|j^E_0(0)|N, M,
\frac{1}{2}\rangle}{\sqrt{4M^2+Q^2}}
\\
\label{eq:em3b}
G^N_M(Q^2)
&=&
\frac{\langle N, \bar{P}, \frac{1}{2}|j^E_+(0)|N, M,
-\frac{1}{2}\rangle}{2\sqrt{Q^2}}
\end{eqnarray}
\end{subequations}
where $\left| N, \bar P, \lambda\right\rangle$ denotes a nucleon state
with four-mo\-men\-tum $\bar P$ and helicity $\lambda$. Furthermore
$Q^2 = -q^2$ and $j^E_\pm := j^E_1 \pm \textrm{i}\,j^E_2$\, as well as
$|\vec P|^2 = Q^2 +$ $\frac{(M^2-{M'}^2-Q^2)^2}{4\,M^2}$\,.
Likewise, the axial vector form factor is given by
\begin{eqnarray}
\label{eq:em4}
G_A(Q^2)
&=&
\frac{\langle p, \bar{P}, \frac{1}{2}|j^A_+(0)|n, M,
-\frac{1}{2}\rangle}{\sqrt{4M^2+Q^2}}
\end{eqnarray}
where again $j^A_\pm := j^A_1 \pm$ $\textrm{i}\,j^A_2$ with $j^A_\mu$ the
axial current operator, whose matrix elements are given by Eq.~(\ref{eq:em1})
after the formal substitution $\hat q \,\mapsto\,$ $\tau^+\gamma^5$\,.
Of course, the normalisations of the form factors is such that the static
magnetic moments and the axial coupling are given by
\begin{displaymath}
\mu_M := G_M(Q^2=0)\,,\qquad g_A := G_A(Q^2=0)\,.
\end{displaymath}
\subsection{The electric form factors of the nucleon\label{nucleon_EFF}}
In fig.~\ref{EFF_Proton} and \ref{EFF_Neutron} we display the electric proton
and neutron form factor, respectively, up to a momentum transfer of $Q^2=6.0$
$\textrm{GeV}^2$. The black solid curve is the result of the present model
$\mathcal{C}$, the blue dashed curve is the result obtained with the
para\-meters of model $\mathcal{A}$, as in~\cite{Merten}, albeit with a better
numerical precision, see the end of Subsection~\ref{subsec:BSAppr}\,.
Although the electric form factor of the proton, see
fig.~\ref{EFF_Proton},
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$G_E^p(Q^2)/G_D(Q^2)$}
\psfrag{MMD}[r][r]{\scriptsize MMD \cite{Mergell}}
\psfrag{Christy}[r][r]{\scriptsize Christy \cite{Christy}}
\psfrag{Qattan}[r][r]{\scriptsize Qattan \cite{Qattan}}
\psfrag{Merten}[r][r]{\scriptsize model $\mathcal{A}$ \cite{Merten}}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=\linewidth]{EFF_Dipole_P.eps}
\caption{
The electric form factor of the proton divided by the dipole form $G_D(Q^2)$, Eq.~(\ref{eq:EFF}).
MMD-Data are taken from Mergell \textit{et al.}~\cite{Mergell}, supplemented by
data from Christy \textit{et al.}~\cite{Christy} and Qattan \textit{et al.}~\cite{Qattan}\,.
The solid black line represents the results from the present model
$\mathcal{C}$; the dashed blue line those from
model $\mathcal{A}$ of~\cite{Merten}, albeit
recalculated with higher numerical precision. Red data points are taken
from polarisation experiments and black ones are obtained by Rosenbluth separation.
\label{EFF_Proton} }
\end{figure}
as calculated with model $\mathcal{A}$ in~\cite{Merten} fell too steeply in
comparison to experimental data, with the present interaction we find a much
improved shape which yields a satisfactory description even up to momentum
transfers of 6~GeV$^2$\,. Indeed, in contrast to model $\mathcal{A}$, which
mainly failed with respect to the isovector part of the form factor, in the
present model $\mathcal{C}$ this form factor shows an almost perfect dipole
shape with the parametrisation
\begin{align}
\label{eq:EFF}
G_D(Q^2) = \frac{1}{(1+Q^2/M_V^2)^2}\,,
\end{align}
taken from \cite{Mergell,Bodek} with $M_V^2=0.71\,\textrm{GeV}^2$.
The resulting electric neutron form factor, see
fig.~\ref{EFF_Neutron},
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$G_E^n(Q^2)$}
\psfrag{MMD}[r][r]{\scriptsize MMD \cite{Mergell}}
\psfrag{Rosenbluth}[r][r]{\scriptsize \cite{Eden,Herberg,Ostrick,Passchier,Schiavilla}}
\psfrag{Polarisation}[r][r]{\scriptsize \cite{Rohe,Golak,Zhu,Madey,Warren,Glazier,Alarcon}}
\psfrag{Merten}[r][r]{\scriptsize model $\mathcal{A}$ \cite{Merten}}
\psfrag{PS-coupl}[r][r]{\scriptsize PS Yukawa}
\psfrag{PV-coupl}[r][r]{\scriptsize PV Yukawa}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=1.0\linewidth]{EFF_N.eps}
\caption{The electric form factor of the neutron. MMD-Data are
taken from the compilation of Mergell \emph{et al.}~\cite{Mergell}.
The solid black line represents the results from the present model
$\mathcal{C}$; the green- and brown-dashed lines corresponds to the
Yukawa-like models with pseudoscalar and/or pseudovector coupling,
see Eqs.~(\ref{eq:OBEVps}) and~(\ref{eq:OBEVpv}) respectively;
the dashed blue line in the result from model $\mathcal{A}$
of~\cite{Merten}, albeit recalculated with higher numerical precision.
Red data points are taken from polarisation experiments and black
ones are obtained by Rosenbluth separation.
\label{EFF_Neutron}}
\end{figure}
has a maximum at approximately the experimental value of $Q^2 \approx 0.4
\textrm{ GeV}^2$ but underestimates the experimental data
from~\cite{Schiavilla} by about the same amount as the earlier calculation
overestimated the data. However, the prediction of model $\mathcal{C}$ is very
similar to the predictions of the Graz group \cite{Plessas} and \cite{Melde07}
for the Goldstone-boson-exchange quark models. The corresponding charge radii
are given in table~\ref{TAB_StatProp}\,. As for the form factor the resulting
squared charge radius of the neutron is calculated too small by a factor of
two. Also the r.m.s. proton radius is slightly smaller than the experimental
value.
\begin{table}[htb]
\centering
\parbox{\linewidth}{
\caption{
Static properties of the nucleon. The values in parentheses are as reported
in~\cite{Merten}, the values on top of these are obtained within the same
model $\mathcal{A}$ but with higher numerical accuracy. The columns PS and
PV are the results obtained with the additional interaction kernels according
to Eqs.~(\ref{eq:OBEVps}) and~(\ref{eq:OBEVpv}), respectively, see also
Section~\ref{sec:Conc} for a brief remark. The static values are extra\-polated
from a dipole shape-like fit.
\label{TAB_StatProp}
}
}
\begin{tabular}{lc@{\hspace*{2pt}}c@{\hspace*{2pt}}c@{\hspace*{2pt}}c@{\hspace*{2pt}}c@{\hspace*{2pt}}c@{\hspace*{2pt}}r}
\toprule
& Model $\mathcal{A}$ & PS & PV & model $\mathcal{C}$ & exp. & ref.\\\midrule
\multirow{2}{*}{$\mu_p[\mu_N]\!\!\!\!$} & 2.76 & \multirow{2}{*}{2.49} & \multirow{2}{*}{2.39} & \multirow{2}{*}{2.54} & \multirow{2}{*}{2.793\!\!\!\!}& \multirow{2}{*}{\cite{PDG}}\\
& [2.74] & & & & & \\\midrule
\multirow{2}{*}{$\mu_n[\mu_N]\!\!\!\!$} & -1.71 & \multirow{2}{*}{-1.59}& \multirow{2}{*}{-1.54}& \multirow{2}{*}{-1.59}& \multirow{2}{*}{-1.913\!\!\!\!}& \multirow{2}{*}{\cite{PDG}}\\
& [-1.70] & & & & \\\midrule
\multirow{2}{*}{$\!\!\!\!\sqrt{\langle r^2\rangle_E^p}$[fm]} & 0.91 & \multirow{2}{*}{0.83} & \multirow{2}{*}{0.68} & \multirow{2}{*}{0.81} & \multirow{2}{*}{0.847\!\!\!\!} & \multirow{2}{*}{\cite{Mergell}}\\
& [0.82] & & & & \\\midrule
\multirow{2}{*}{$\langle r^2\rangle_E^n$[fm]$^2$} & -0.20 & \multirow{2}{*}{0.01}& \multirow{2}{*}{0.08} & \multirow{2}{*}{-0.06}& -0.123\!\!\!\! & \multirow{2}{*}{\cite{Mergell}}\\
& [-0.11] & & & & $\pm$0.004 & \\\midrule
\multirow{2}{*}{$\!\!\!\!\sqrt{\langle r^2\rangle_M^p}$[fm]} & 0.90 & \multirow{2}{*}{0.81} & \multirow{2}{*}{0.68} & \multirow{2}{*}{0.78} & \multirow{2}{*}{0.836\!\!\!\!} & \multirow{2}{*}{\cite{Mergell}}\\
& [0.91] & & & & & \\\midrule
\multirow{2}{*}{$\!\!\!\!\sqrt{\langle r^2\rangle_M^n}$[fm]} & 0.84 & \multirow{2}{*}{0.79} & \multirow{2}{*}{0.67} & \multirow{2}{*}{0.75} & \multirow{2}{*}{0.889}\!\!\!\! & \multirow{2}{*}{\cite{Mergell}}\\
& [0.86] & & & & & \\\midrule
\multirow{2}{*}{$g_A$} & 1.22 & \multirow{2}{*}{1.17} & \multirow{2}{*}{1.14} & \multirow{2}{*}{1.13} & 1.267\!\!\!\! & \multirow{2}{*}{\cite{PDG,Bodek}}\\
& [1.21] & & & & $\pm$0.0035\!\!\!\! & \\\midrule
\multirow{2}{*}{$\!\!\sqrt{\langle r^2\rangle_A}$[fm]} & 0.68 & \multirow{2}{*}{0.64} & \multirow{2}{*}{0.48} & \multirow{2}{*}{0.57} & 0.67\!\!\!\! & \multirow{2}{*}{\cite{Bernard}}\\
& [0.62] & & & &$\pm$0.01\!\!\!\! & \\\bottomrule
\end{tabular}
\end{table}
\begin{table}[htb]
\centering
\caption{Octet hyperon magnetic moments $\mu$ for model $\mathcal{A}$ and $\mathcal{C}$
calculated as in \cite{Haupt}. The values are given in units of $\mu_N$.\label{TAB_Octet}}
\begin{tabular}{c@{\hspace*{10pt}}c@{\hspace*{10pt}}c@{\hspace*{10pt}}c@{\hspace*{10pt}}}
\toprule
hyperon & model $\mathcal{A}$ & model $\mathcal{C}$ & PDG \cite{PDG} \\\midrule
$\Lambda$ & -0.606 & -0.577 & -0.613 $\pm$0.004\\\midrule
$\Sigma^+$ & 2.510 & 2.309 & 2.458 $\pm$0.010\\\midrule
$\Sigma^0$ & 0.743 & 0.701 & - \\\midrule
$\Sigma^-$ & -1.013 & -0.908 & -1.160 $\pm$0.025\\\midrule
$\Xi^0$ & -1.324 & -1.240 & -1.250 $\pm$0.014\\\midrule
$\Xi^-$ & -0.533 & -0.532 & -0.651$\pm$0.0025\\\bottomrule
\end{tabular}
\end{table}
\begin{table}[htb]
\centering
\caption{Decuplet hyperon magnetic moments $\mu$ for model $\mathcal{A}$ and $\mathcal{C}$
calculated as in \cite{Haupt}. The values are given in units of $\mu_N$.\label{TAB_Decuplet}}
\begin{tabular}{c@{\hspace*{10pt}}c@{\hspace*{10pt}}c@{\hspace*{10pt}}c@{\hspace*{10pt}}}
\toprule
hyperon & model $\mathcal{A}$ & model $\mathcal{C}$ & PDG \cite{PDG} \\\midrule
$\Delta^{++}$ & 4.241 & 4.238 & 3.7 to 7.5\\\midrule
$\Delta^+$ & 2.121 & 2.119 & $2.7_{-1.3}^{+1.0}\pm1.5\pm3$\\\midrule
$\Delta^0$ & 0.0 & 0.0 & \\\midrule
$\Delta^-$ &-2.121 &-2.119 & \\\midrule
$\Sigma^{\ast+}$ & 2.567 & 2.431 & \\\midrule
$\Sigma^{\ast0}$ & 0.275 & 0.205 & \\\midrule
$\Sigma^{\ast-}$ &-2.017 &-2.021 & \\\midrule
$\Xi^{\ast0}$ & 0.607 & 0.474 & \\\midrule
$\Xi^{\ast-}$ &-1.865 &-1.765 & \\\midrule
$\Omega^-$ & -1.675& -1.577 & $-2.02\pm0.05$\\\bottomrule
\end{tabular}
\end{table}
In fig.~\ref{MFF_Proton} and ~\ref{MFF_Neutron} we display the magnetic
proton- and neutron form factor up to a momentum transfer of $Q^2=6.0\,
\textrm{GeV}^2$\,, respectively.
Again, the black solid curve is the result of the present model $\mathcal{C}$, the blue dashed
curve is the result obtained with the parameters of model $\mathcal{A}$, as
in~\cite{Merten}, albeit with a better numerical precision, see the end of
Subsection~\ref{subsec:BSAppr}\,.
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$G_M^p(Q^2)/G_D(Q^2)/\mu_p$}
\psfrag{MMD}[r][r]{\scriptsize MMD \cite{Mergell}}
\psfrag{Christy}[r][r]{\scriptsize Christy \cite{Christy}}
\psfrag{Qattan}[r][r]{\scriptsize Qattan \cite{Qattan}}
\psfrag{Bartel}[r][r]{\scriptsize Bartel \cite{Bartel}}
\psfrag{Merten}[r][r]{\scriptsize Merten $\mathcal{A}$ \cite{Merten}}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=1.0\linewidth]{MFF_Dipole_P.eps}
\caption{The magnetic form factor of the proton divided by the dipole form
$G_D(Q^2)$, Eq.~(\ref{eq:EFF}) and the magnetic moment of the
proton $\mu_p=2.793\,\mu_N$. MMD-Data are taken from the compilation
of Mergell \emph{et al.}~\cite{Mergell}. Additionally, polarisation
experiments are marked by red. The black marked data points are obtained by
Rosenbluth separation.\label{MFF_Proton}}
\end{figure}
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$G_M^n(Q^2)/G_D(Q^2)/\mu_n$}
\psfrag{MMD}[r][r]{\scriptsize MMD \cite{Mergell}}
\psfrag{Anklin}[r][r]{\scriptsize Anklin \cite{Anklin}}
\psfrag{Kubon}[r][r]{\scriptsize Kubon \cite{Kubon}}
\psfrag{Xu}[r][r]{\scriptsize Xu \cite{Xu}}
\psfrag{Madey}[r][r]{\scriptsize Madey \cite{Madey}}
\psfrag{Alarcon}[r][r]{\scriptsize Alarcon \cite{Alarcon}}
\psfrag{Merten}[r][r]{\scriptsize model $\mathcal{A}$ \cite{Merten}}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=1.0\linewidth]{MFF_Dipole_N.eps}
\caption{The magnetic form factor of the neutron divided by the dipole form $G_D(Q^2)$,
Eq.~(\ref{eq:EFF}) and the magnetic moment of the neutron $\mu_n=-1.913\,\mu_N$.
MMD-Data are taken from the compilation by Mergell \emph{et al.}~\cite{Mergell} and
from more recent results from MAMI~\cite{Anklin,Kubon}\,. Additionally, polarisation
experiments are marked by red data points. The black marked ones are obtained by
Rosenbluth separation.\label{MFF_Neutron}}
\end{figure}
Whereas in the original calculation (model $\mathcal{A}$ of~\cite{Merten}) the
absolute value of these form factors dropped slightly too fast as a function
of the momentum transfer, in the present calculation we now find a very good
description even at the highest momentum transfers. Only at low momentum
transfer the values are too small as is reflected by the rather small values
for the various magnetic radii, see table~\ref{TAB_StatProp} and the too small
values of the calculated magnetic moments. Note, however that the ratio
$\mu_p/\mu_n \approx 1.597$ for model $\mathcal{C}$ slightly changes (previously
$\mu_p/\mu_n \approx 1.605$ for model $\mathcal{A}$) and is slightly larger
than the experimental value $\mu_p/\mu_n \approx 1.46$\,; all values are
remarkably close to the non-relativistic constituent quark model value
$\mu_p/\mu_n = \frac{3}{2}$\,. The magnetic moments of flavour octet and decuplet
baryons has been calculated accordingly to the method outlined in \cite{Haupt}. The
results are compared to experimental values in Table \ref{TAB_Octet} and
\ref{TAB_Decuplet}, respectively. As a consequence of the better description
of the momentum transfer dependencies in the individual form factors, we now also
find an improved description of the momentum transfer dependence of the form factor
ratio $\mu_p\,G^p_E / G^p_M(Q^2)$\,, which has been the focus on the discussion
whether two-photon amplitudes are relevant for the discrepancy~\cite{Vanderhaeghen}
found between recent measurements based on polarisation data (red data points of
fig.~\ref{GEpGMp})~\cite{Milbrath,Jones,Gayou2001,Gayou2002,Punjabi,Hu,Crawford,Higinbotham,Ron,Zhan}
versus the traditional Rosenbluth separation (black data points of fig.~\ref{GEpGMp}),
see \emph{e.g.}~\cite{Price,Bartel,Berger_1,Walker}. Whereas in
the original model $\mathcal{A}$ this ratio fell much too steep, we now find
in model $\mathcal{C}$ a much better description of this quantity, see
fig.~\ref{GEpGMp} for a comparison with various data. Up to $Q^2 \approx 3\,\textrm{GeV}^2$
we indeed find the observed linear dependence.
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$\mu_p\,G_E^p / G_M^p(Q^2)$}
\psfrag{Rosenbluth}[r][r]{\scriptsize \cite{Price,Bartel,Berger_1,Walker,Andivahis}}
\psfrag{Polarisation}[r][r]{\scriptsize \cite{Milbrath,Jones,Gayou2001,Pospischil,Gayou2002,Punjabi,Higinbotham,Hu,Crawford,Ron,Zhan}}
\psfrag{Merten}[r][r]{\scriptsize model $\mathcal{A}$ \cite{Merten}}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=1.0\linewidth]{E_MFF_P_ratio.eps}
\caption{The ratio $\mu_p\,G_E^p/G_M^p$ compared to recent JLAB data (see
legend).
In the insert the low momentum transfer region is enlarged. The solid
black line is the present result, the dashed blue line the
result in model $\mathcal{A}$\,. Red data points are taken from polarisation
experiments and the black ones are obtained from Rosenbluth separation.
\label{GEpGMp}}
\end{figure}
Finally, the axial form factor, see fig.~\ref{AFF_3}, was already very well
described in model $\mathcal{A}$ of~\cite{Merten}\,. Although falling slightly
less steeply, the present calculation still gives a very satisfactory
description of the data also at higher momentum transfers in the same manner
as in~\cite{Glozman2001,Wagenbrunn2001,Wagenbrunn2003}. As for the magnetic
moments the value of the axial coupling constant is too small, but of course
much better than the non-relativistic constituent quark model result $g_A=\frac{5}{3}$\,.
The axial form factor, presented in fig.~\ref{AFF_3}, is divided by the axial dipole form
\begin{align}
\label{eq:AFF}
G^A_D(Q^2)=\frac{g_A}{(1+Q^2/M_A^2)^2}\,,
\end{align}
with the parameters $M_A=1.014\pm0.014\,\textrm{GeV}$
and $g_A=1.267$ taken from Bodek \textit{et al.} \cite{Bodek}.
\begin{figure}[!htb]
\centering
\psfrag{x-axis}[c][c]{$Q^2\,\,[\textrm{GeV}^2]$}
\psfrag{y-axis}[c][c]{$G_A^3(Q^2)/G_D^A(Q^2)/g_A$}
\psfrag{Amaldi}[r][r]{\scriptsize Amaldi \cite{Amaldi}}
\psfrag{Brauel}[r][r]{\scriptsize Brauel \cite{Brauel}}
\psfrag{Bloom}[r][r]{\scriptsize Bloom \cite{Bloom}}
\psfrag{Del Guerra}[r][r]{\scriptsize Del Guerra \cite{DGuerra}}
\psfrag{Joos}[r][r]{\scriptsize Joos \cite{Joos}}
\psfrag{Baker}[r][r]{\scriptsize Baker \cite{Baker}}
\psfrag{Miller}[r][r]{\scriptsize Miller \cite{Miller}}
\psfrag{Kitagaki83}[r][r]{\scriptsize Kitagaki 83 \cite{Kitagaki83}}
\psfrag{Kitagaki90}[r][r]{\scriptsize Kitagaki 90 \cite{Kitagaki90}}
\psfrag{Allasia}[r][r]{\scriptsize Allasia \cite{Allasia}}
\psfrag{Merten}[r][r]{\scriptsize model $\mathcal{A}$ \cite{Merten}}
\psfrag{Gauss}[r][r]{\scriptsize model $\mathcal{C}$}
\includegraphics[width=1.0\linewidth]{AFF_Dipole_3.eps}
\caption{The axial form factor of the nucleon divided by the axial dipole
form in Eq.~(\ref{eq:AFF}) and the axial coupling $g_A=-1.267$. The black solid
line is the present result, the blue dashed line the
result in model $\mathcal{A}$\,. Experimental
data are taken from the compilation by Bernard \emph{et al.}~\cite{Bernard}\,
\label{AFF_3}\,.}
\end{figure}
In summary we find, that the new model $\mathcal{C}$, apart from some
improvements in the description of the excitation\,\,\,\, spec\-tra at the expense of
additional parameters of a phenomenologically introduced flavour dependent
interaction does allow for a parameter-free description of electromagnetic
gr\-ound state properties of a similar overall quality as has been obtained
before, with some distinctive improvements on the momentum transfer dependence
of various form factors. A discussion of the momentum dependence of helicity
amplitudes for the electromagnetic excitation of baryon resonances will be
given in a subsequent paper~\cite{Ronniger2}\,.
\section{Summary and conclusion}
\label{sec:Conc}
In the present paper we have tried to demonstrate that by introducing an
additional flavour dependent interaction, para\-metrised with a Gaussian radial
dependence with an universal range and two couplings for flavour octet and
flavour singlet exchange, it is possible to improve upon some deficiencies
found in a former relativistically covariant constituent quark model treatment
of baryonic excitation spectra based on (an instantaneous formulation of) the
Bethe-Salpeter equation. These improvements include:
\begin{itemize}
\item
A better description of excited negative parity states sli\-ght\-ly below $2$
GeV in the $\Delta$ spectrum;
\item
A better description of the position of the first scalar, iso\-scalar
excitation of the ground state in all light-fla\-vour sectors;
\item
An improved description of the momentum de\-pend\-ence of electromagnetic form
factors of ground states without the introduction of any additional
parameters.
\end{itemize}
It must be conceded that this additional interaction was introduced purely
phenomenologically and required a\linebreak drastic modification of the parametrisation
of confinement and the other flavour dependent interaction of the original model, which
had a form as inferred from instanton effects. In spite of this, with only 10
parameters in total we still consider this to be an effective description of
the multitude of resonances found for baryons made out of light flavoured
quarks.
\begin{figure*}[!htb]
\centering
\psfrag{De}[c][c]{\Huge$\Delta$}
\psfrag{Nu}[c][c]{\Huge$N$}
\includegraphics[width=0.8\linewidth]{Summary_Spec.eps}
\parbox{\linewidth}{
\caption{
Comparison of the $\Delta$-Spectrum (first six columns) and the
$N$-Spectrum (rightmost six columns) calculated within the PS and PV coupled models
from Eqs.~(\ref{eq:OBEVps}) and~(\ref{eq:OBEVpv}) (right and left side of
each column) and with the experimental data from the Particle Data Group~\cite{PDG}
(central in each column).
\label{Summary_Spec}
}
}
\end{figure*}
Nevertheless, it would have been preferred, if the additional flavour dependent
interaction could be related to a genuine physical process, such as light
pseudoscalar meson exchange. Indeed, we tried to find parametrisations of
confinement and parameters of the instanton induced interaction, that could be
combined with interaction kernels as given by the expressions in
Eq.~(\ref{eq:OBEVps}) or~(\ref{eq:OBEVpv}). However, for pseudoscalar
coupling (PS) of a meson nonet to the quarks, we could find a description of the
mass spectra with similar features and of a similar quality as in the present
model $\mathcal{C}$ only at the expense of introducing flavour SU(3) symmetry
breaking and thus introducing more parameters. The latter was also found to be
the case for pseudovector coupling (PV) where, moreover, no significant improvement
concerning the deficiencies in the spectra mentioned above was
found. For the sake of completeness, the nucleon spectrum for PS and PV coupled models
from Eqs.~(\ref{eq:OBEVps}) and~(\ref{eq:OBEVpv}) is shown in fig.~\ref{Summary_Spec}.
Furthermore, with both \emph{An\-satze} we were not able to reproduce in
particular the electric neutron form factor, see fig.~\ref{EFF_Neutron} and
table~\ref{TAB_StatProp} for some typical results. Accordingly, we dismissed
these possibilities and preferred the phenomenological approach of model
$\mathcal{C}$ dis\-cus\-sed in the paper.
A parameter-free calculation of longitudinal and transverse helicity amplitudes
for electro-excitation is presently performed and the results will be dis\-cussed
in a subsequent paper~\cite{Ronniger2}\,.
|
\section{Introduction}\label{sec|intro}
Supermassive black holes reside at the center of most, if not all,
massive galaxies. The masses of black holes are tightly
correlated with properties of their galactic hosts, especially
the velocity dispersions and masses of their stellar bulges
(e.g., Ferrarese \& Merritt 2000; Gebhardt et al.\ 2000;
H\"aring \& Rix 2004). If black holes grow principally
by radiatively efficient accretion, then the statistics
of quasars and active galactic nuclei (AGN) can be used to
track the growth of the black hole population.
One common approach to modeling this growth uses the
black hole ``continuity equation'' (Cavaliere et al.\ 1971;
Small \& Blandford 1992). Simple versions of such models
have two free parameters: the radiative efficiency $\epsilon$,
which relates mass accretion rate to luminosity, and the Eddington
ratio $\lambda = L/L_{\rm Edd}$, which relates luminosity
to black hole mass. With the simplifying assumption that all active
black holes have the same fixed values of $\epsilon$ and $\lambda$, and
that each black hole has a duty cycle (i.e., a
probability of being active at a given time)
that depends on mass and redshift, $U(M_{\rm BH},z)$ ,
one can use the observed AGN luminosity function to infer the
average rate at which mass is being added to each mass range
of the black hole population, thus evolving the black hole mass
function forward in time. Model assumptions can be tested
by comparing to observational estimates of the local ($z=0$)
black hole mass function (e.g., Salucci et al. 1998; Marconi et al.\ 2004;
Shankar et al.\ 2004; Shankar, Weinberg \& Miralda-Escud\'e 2009, hereafter
SWM).
The models presented in SWM made use of this simplifying assumption of
fixed values of $\epsilon$ and $\lambda$. Our key findings were
{\it (i)} that models with $\epsilon \approx 0.07$ and
$\lambda \approx 0.25$ yield a reasonable match to the
local black hole mass function, though these specific
inferred parameter values are sensitive to remaining uncertainties in the
bolometric
AGN luminosity function, and
{\it (ii)} that reproducing the observed luminosity function with these models
required ``downsizing'' evolution, with the duty
cycle for high mass black holes declining more rapidly with time than the
duty cycle for low mass black holes.
Additional constraints on duty cycles (apart from their relation to the
fraction of active galaxies) can be inferred from AGN and
quasar clustering
(Haiman \& Hui 2001; Martini \& Weinberg 2001), although these
constraints depend on the assumed level of scatter
between AGN luminosity and the mass of the host dark matter
halo. Reproducing the strong observed clustering of
$z=4$ quasars (Shen et al.\ 2007) requires duty cycles
close to one and minimal scatter between luminosity and
halo mass (White, Martini \& Cohn 2008; Shankar et al. 2010b).
However, reconciling the duty cycles predicted by SWM with the
lower redshift clustering ($z\approx 1.5$) measured by
Shen et al.\ (2009) requires significant scatter
($\Sigma \approx 0.5\,$dex) between luminosity and halo
mass to lower the predicted clustering amplitude
(Shankar, Weinberg \& Shen 2010c).
In this paper we consider more general continuity-equation models than
those in SWM, incorporating a distribution of Eddington ratios $P(\lambda)$,
as well as
allowing for possible redshift evolution or black hole mass dependence
of characteristic $\lambda$ values or average radiative efficiency.
The motivation for these more general models is to make contact with
a broader range of observations that offer additional constraints on
AGN and black hole evolution. Most obviously, many authors have
now used black hole masses inferred from linewidth measurements
or host galaxy properties to directly estimate Eddington ratios,
finding evidence for broad $P(\lambda)$\ distributions that change with
redshift and black hole mass (see Section~\ref{subsec|CompareWithObservedPL}
below for observational references). Active galaxy fractions provide
another set of important constraints on models. In single-$\lambda$
models the duty cycle $U(M_{\rm BH},z)$\ --- the fraction of black holes of mass $M_{\rm BH}$
at redshift $z$ that are active at a given time ---
follows simply from the ratio of the observed luminosity function
to the calculated black hole mass function (see SWM).
With a broad $P(\lambda)$, the definition of duty cycle depends on the
adopted threshold in luminosity or Eddington ratio, and because
observational samples inevitably include these thresholds, the observed
active galaxy fraction also tests models of $P(\lambda)$.
Other authors have recently implemented continuity equation models
with broad Eddington ratio distributions
(Merloni \& Heinz 2008; Yu \& Lu 2008; Cao 2010),
but their studies are limited to a few specific models.
Our implementation here draws on some of the techniques introduced
by Cao (2010) and also on techniques and ideas from Steed \& Weinberg (2003),
who attempted a broad-ranging study of black hole evolution within a
general continuity equation framework.
The broad $P(\lambda)$\ models described here also provide a more
general framework for modeling quasar and AGN clustering,
a topic we will examine in future work.
Phenomenological models of the black hole population like
those in this paper complement models based on hydrodynamic
simulations (e.g., di Matteo et al.\ 2005) or
semi-analytic models that tie quasar activity to major
mergers of galaxies or dark matter halos
(e.g., Kauffmann \& Haehnelt 2000; Wyithe \& Loeb 2003;
Hopkins et al. 2006). The numerical and semi-analytic
approaches adopt a more complete physical scenario for the
mechanisms that drive black hole accretion, and they tie
AGN to the underlying galaxy population by construction.
However, it can be difficult to interpret the significance
of discrepancies with observations, or to know whether observational
successes imply that the physical assumptions are correct.
Phenomenological models are free to be driven by the
data, within the constraints imposed by the adopted parameterization.
The next section describes our methods for computing black
hole evolution and duty cycles, and it summarizes the SWM
estimate of the bolometric luminosity function that is our
basic observational input.
All of our models reproduce this luminosity function by construction.
In Section~\ref{sec|broadPL} we
describe our three basic models for $P(\lambda)$\ --- a
$\delta$-function, a Gaussian in $\log\lambda$, and a
Gaussian plus a power-law tail to low $\lambda$ ---
then examine their impact on mass function evolution and
conduct a first comparison to observed $\lambda$ distributions.
In Section~\ref{sec|ActiveFractions} we turn our focus to
active galaxy fractions and discuss how data might favour
redshift evolution and possibly mass dependence of $P(\lambda)$.
In Section~\ref{sec|OtherConstraints}
we revisit the comparison to observed $\lambda$ distributions,
the impact of black hole mergers at the rate predicted by
hierarchical galaxy formation models, the relation between specific
black hole growth and specific star formation rate,
and the impact of adopting a $\lambda$-dependent bolometric correction.
In Section~\ref{sec|discu} we discuss our results highlighting some internal tensions
between data sets and key tensions between models and data.
We conclude in Section~\ref{sec|conclu}.
The appendices present some details of our calculational methods
and of our estimates of observed active galaxy fractions.
Throughout the paper we use cosmological parameters $\Omega_m=0.30$,
$\Omega_\Lambda=0.70$, and $h\equiv H_0/100\, {\rm km\, s^{-1}\, Mpc^{-1}}=0.7$,
and we compute the linear matter power spectra running the
Smith et al.\ (2003) code with $\Gamma=0.19$, and $\sigma_8=0.8$.
\section{Formalism}\label{sec|formalism}
\subsection{Evolving the black hole mass function}
\label{subsec|BHMFevol}
To study the global evolution of the black hole population through
time, we develop models in which the black hole mass function is
self-consistently evolved via the continuity equation,
\begin{equation}
\frac{\partial n_{\rm BH}}{\partial
t}(M_{\rm BH},t)=-\frac{\partial (\langle \dot{M}_{\rm BH}\rangle
n_{\rm BH}(M_{\rm BH},t))}{\partial M_{\rm BH}}\,,
\label{eq|conteq}
\end{equation}
where $\langle \dot{M}_{\rm BH}\rangle$ is the mean accretion
rate (averaged over the active and inactive populations) of the
black holes of mass $M_{\rm BH}$\ at time $t$
(see, e.g.,
Cavaliere et al. 1971; Small \& Blandford 1992; Yu \& Tremaine 2002; Steed
\& Weinberg 2003; Marconi et al. 2004; Merloni 2004; Yu \& Lu 2004; Tamura
et al. 2006;
SWM; Shankar 2009; Raimundo \& Fabian 2009; Kisaka \& Kojima 2010).
This evolution is equivalent
to the case in which every black hole grows constantly at the mean
accretion rate $\langle \dot{M}_{\rm BH}\rangle$. In practice, individual
black holes turn on and off, but the mass function evolution depends only on
the mean accretion rate as a function of mass. Note that Eq.~(\ref{eq|conteq})
neglects any contribution from black hole
mergers, which do not add mass to the black hole population but can alter
the mass function by redistribution.
In Section~\ref{subsec|Mergers} we discuss the impact
of including black hole mergers at the rate predicted by hierarchical models
of structure formation.
We first define the Eddington ratio,
\begin{equation}
\lambda \equiv {L \over L_{\rm Edd}},
\end{equation}
where
\begin{equation}
L_{\rm Edd}\equiv l M_{\rm BH} =1.26\times 10^{38}\left(\frac{M_{\rm BH}}{{\rm
M_{\odot}}}\right) \, {\rm erg\, s^{-1}}\,
\label{eq|Lbol}
\end{equation}
is the Eddington (1922) luminosity computed for Thomson scattering
opacity and pure hydrogen composition. The growth rate of an active
black hole of mass $M_{\rm BH}$\ with Eddington ratio $\lambda$\ is $\dot{M}_{\rm
BH}=$$M_{\rm BH}$$/t_{\rm ef}$,
where the \emph{e}-folding time is (Salpeter 1964)
\begin{equation}
t_{\rm ef}=4\times 10^8\left(\frac{f}{\lambda}\right)\, {\rm yr}
\equiv {t_s \over \lambda}
~,
\label{eq|tefold}
\end{equation}
with
\begin{equation}
f=\epsilon/(1-\epsilon) \quad \hbox{and} \quad
\epsilon=\frac{L}{\dot{M}_{\rm inflow}c^2}. \label{eq|epsilon}
\end{equation}
The radiative efficiency $\epsilon$ is conventionally defined with
respect to the large scale mass inflow rate, but the black hole
growth rate $\dot{M}_{\rm BH}$ is smaller by a factor of $(1-\epsilon)$
because of radiative losses.
We then define the probability for a black hole of a given mass
$M_{\rm BH}$\ to accrete at the Eddington ratio $\lambda$\,
per unit $\log\lambda$, at redshift $z$ as
$P(\lambda|M_{\rm BH},z)$, normalized to unity, i.e.,
$\int d\log \lambda \, P(\lambda|M_{\rm BH},z)=1$.\footnote{Throughout
the paper, log denotes base-10 logarithm and ln denotes natural logarithm.}
This is the
quantity that can be related to the observed Eddington ratio distributions
discussed in Section~\ref{sec|intro}.
The average growth
rate of all black holes can be computed by
convolving the input $P(\lambda|M_{\rm BH},z)$\ with the duty cycle $U(M_{\rm BH},z)$,
\begin{equation}
\langle \dot{M}_{\rm BH}\rangle=\int d\log \lambda \, P(\lambda|M_{\rm
BH},z)\lambda \, U(M_{\rm BH},z)\, \frac{M_{\rm BH}}{t_{\rm s}}\, ,
\label{eq|MdotAve}
\end{equation}
where the integral extends over all allowed values of $\lambda$.
A physically consistent model must have $U(\lambda,M_{\rm BH},z)\le 1$
for all $M_{\rm BH}$\ and $z$.
In models where the Eddington ratio has a single value $\lambda_1$,
$P(\lambda|M_{\rm BH},z)=\lambda_1 \delta(\lambda - \lambda_1)$,
the duty cycle is simply the ratio of the luminosity and black hole mass
functions,\footnote{Following SWM, we will always use the symbol $\Phi(x)$
to denote mass and luminosity
functions in logarithmic units of $L$ or $M_{\rm BH}$,
i.e. $\Phi(x)d\log x=n(x)x \ln(10)dx$,
where $n(x)dx$ is the comoving space density of black holes in the
mass or luminosity range $x\rightarrow x+dx$, in units of ${\rm Mpc^{-3}}$
for $h=0.7$.
Unless otherwise stated, all masses are in units of $M_\odot$ and all
luminosities are bolometric
and in units of ${\rm erg\, s^{-1}}$; e.g., $\log L > 45$ refers to quasars with bolometric
luminosity $L > 10^{45} {\rm erg \, s^{-1}}$.}
\begin{equation}
U(M_{\rm BH},z)=\frac{\Phi(L,z)}{\Phi_{\rm BH}(M_{\rm BH},z)}\, ,
\qquad\qquad L=\lambda l M_{\rm BH}\, .
\label{eq|P0general}
\end{equation}
\begin{figure*}
\includegraphics[width=15truecm]{Figure1.eps}
\caption{\emph{Left}: Comparison among the three $P(\lambda)$\ distributions
taken as a reference throughout the paper.
The \emph{dotted} line is the $\delta$-model centered on $\log \lambda_c$$=-0.6$, the \emph{solid} line is
the G-model, a Gaussian with dispersion of $\Sigma_{\lambda}$$=0.3$ dex and centered around
the same value of $\log \lambda_c$$=-0.6$, and
the \emph{long-dashed} line is the G+P-model, characterized by the same Gaussian plus
a power-law with slope $\alpha=-0.3$ normalized
to have the same value as the Gaussian at $\log \lambda=\log \lambda_c-
\log \Sigma_{\lambda}$. Note that
we define our distributions in $\log \lambda$ but refer to them in
figures and text as $P(\lambda)$,
and that all distributions are truncated at $\lambda=1$. \emph{Right}:
Dimensionless emissivity per $\log\lambda$ interval,
$\lambda P(\lambda)$, for the G and G+P distributions.
Note that even the G+P model has its emissivity peak near
the peak of the Gaussian component.}
\label{fig|comparePLambda}
\end{figure*}
This paper presents predictions from several models that take as
input the observed luminosity function $\Phi(L,z)$, an assumed
Eddington ratio distribution of active black holes, $P(\lambda|M_{\rm BH},z)$, and an
assumed radiative efficiency $\epsilon$, to solve
Eq.~(\ref{eq|conteq}) forwards in time. At each timestep,
we use the black hole mass function $\Phi_{\rm BH}(M_{\rm BH},z)$ derived
from the previous timestep (or from an assumed initial
condition in the first timestep), and
we compute the duty cycle $U(M_{\rm BH},z)$\ by numerically solving the equation
\begin{equation}
\Phi(L,z)=\int d\log \lambda P(\lambda|M_{\rm BH},z) U(M_{\rm BH},z)
\Phi_{\rm BH}(M_{\rm BH},z)\,
\label{eq|PhiLLambda}
\end{equation}
with $M_{\rm BH} = L/(\lambda l)$.
We then compute the average growth rate as a function of black hole mass via
Eq.~(\ref{eq|MdotAve}) and, finally,
update the black hole masses and black hole mass function via the
continuity equation (Eq.~\ref{eq|conteq}).
Details of our numerical implementation are given in
Appendix~\ref{app|solvingContEq}.
Directly solving Eq.~(\ref{eq|PhiLLambda}) for $U(M_{\rm
BH},z)$ is feasible for a discrete distribution of Eddington ratios.
However, for all of the models discussed
in this work, we will adopt \emph{continuous} $P(\lambda|M_{\rm BH},z)$\ distributions,
which can be more efficiently handled following the method
described by Steed \& Weinberg (2003) and Cao (2010).
The latter is based on solving
Eq.~(\ref{eq|PhiLLambda}) for the full mass function of \emph{active}
black holes
\begin{equation}
N_{\rm act}(M_{\rm BH},z)=\Phi_{\rm BH}(M_{\rm BH},z) \times U(M_{\rm BH},z) \, .
\label{eq|NactDef}
\end{equation}
More specifically,
at any redshift $z$ a parameterized active mass function $N_{\rm act}(M_{\rm BH},z)$\ is derived by
directly fitting\footnote{As discussed in Appendix~\ref{app|solvingContEq},
we adopt a double power-law form for the input $N_{\rm act}(M_{\rm BH},z)$, which is a good
(though not perfect) approximation of the functional form adopted for the
AGN luminosity function
(see Section~\ref{subsec|AGNLF}).
Our procedure allows us to reproduce the AGN
luminosity function at the $\sim 5\%$ level,
adequate precision given the statistical uncertainties
in the data.}
the AGN luminosity function $\Phi(L,z)$ at the same redshift via
Eq.~(\ref{eq|PhiLLambda}). The function
$N_{\rm act}(M_{\rm BH},z)$\ is then inserted in Eq.~(\ref{eq|conteq}) by replacing
$\langle \dot{M}_{\rm BH}\rangle
n_{\rm BH}(M_{\rm BH},t)=\int d\log \lambda P(\lambda|M_{\rm BH},z)
M_{\rm BH} N_{\rm act}(M_{\rm BH},z)/t_{\rm s}$,
thus allowing the computation of the full black hole mass function $\Phi_{\rm BH}(M_{\rm BH},z)$\ and
the duty cycle $U(M_{\rm BH},z)=N_{\rm act}(M_{\rm BH},z)/\Phi_{\rm BH}(M_{\rm BH},z)$.
Eq.~(\ref{eq|conteq}) is solved by imposing an initial condition
that we fix, as in SWM, by assuming a constant value of
the initial duty cycle.
We usually choose this to be 0.5 at $z=6$ for all masses,
but in some models we lower it to
0.1 to allow the duty cycles at later times
to be always lower than unity.
The results at later redshifts are insensitive to the initial
conditions for mass bins that have experienced substantial accretion
growth (Marconi et al.\ 2004; SWM).
Note that Eq.~(\ref{eq|conteq}) could be solved backwards
in time as well by forcing the initial condition to be the
local black hole mass function (e.g., Merloni 2004).
In this case, incorrect assumptions about the physical parameters
or present day mass function manifest themselves as unphysical
mass functions at earlier redshifts (e.g., negative space densities).
Given the still substantial uncertainties in the local mass function
(e.g., SWM; Vika et al.\ 2009; Shankar et al., in prep.) and our
desire to explore a wide range of physical models, we prefer to
evolve forward in time and compare to the local mass function
as one of several observational constraints.
\subsection{Input AGN Luminosity Function}\label{subsec|AGNLF}
A full discussion of our adopted AGN luminosity function appears
in SWM; here we provide a brief summary.\footnote{The
full tables of the input
AGN luminosity functions, and also black hole mass functions and duty cycles (for a set of
representative models), can be found
in electronic format at {\tt http://mygepi.obspm.fr/$\sim$fshankar/}.}
We adopt the Ueda et al. (2003) fit to
the number of sources per unit volume per dex of luminosity $\log
L_X=\log L_{2-10\, {\rm keV}}$, composed of a
double power-law and an evolution term:
\begin{equation}
\Phi(L_X,z) = e(z,L_X)\, \frac{A}{
\left( \frac{L_X}{L^*}\right)^{\gamma_1} +
\left( \frac{L_X}{L^*}\right)^{\gamma_2} } \, ,
\label{eq|U03}
\end{equation}
where
\begin{equation}
e(L_X,z)=\left\{
\begin{array}{ll}
(1+z)^{p_1} & \hbox{if $z<z_c(L_X)$} \\
(1+z_c)^{p_1}[\frac{1+z}{1+z_c(L_X)}]^{p_2} & \hbox{if $z\ge z_c(L_X)$\,
,}
\end{array}
\right.
\label{eq|eLz}
\end{equation}
with
\begin{equation}
z_c(L_X)=\left\{
\begin{array}{ll}
z_c^* & \hbox{if $L_X\ge L_a$} \\
z_c^*(L_X/L_a)^{0.335} & \hbox{if $L_X< L_a$\, .}
\end{array}
\right.
\label{eq|zc}
\end{equation}
The values of the parameters have been tuned
to correctly match the full set of data presented by SWM (see their Table~1).
The X-ray luminosity function of Eq.~(\ref{eq|U03}) is converted into a bolometric
luminosity function using the fit to the bolometric correction given by
Marconi et
al. (2004), $\log L/L_X=1.54+0.24\zeta+0.012\zeta^2-0.0015\zeta^3$,
with $\zeta=\log L/L_{\odot}-12$, and $L_{\odot}=4\times 10^{33}\
{\rm erg\, s^{-1}}$. We have changed
our procedure slightly with respect to SWM, dropping the convolution
with 0.2-dex Gaussian scatter in bolometric correction, because the
assumption that this scatter is uncorrelated with X-ray luminosity
seems insecure. However, this change makes minimal difference to
the derived luminosity function, and our results using the original
SWM fit would not be noticeably different.
The luminosity function data that we fit also include
a number density of Compton-thick sources in each luminosity bin equal to the number of sources in the column
density range $23\le \log N_H/{\rm cm^{-2}}\le 24$, computed following
the Ueda et al. (2003) prescriptions for the $P(N_H|L)$ column density
distributions. We checked that our estimates of the Compton-thick number densities at high redshifts
are fully consistent with the recent results by Alexander et al. (2011) and Fiore et al. (2011).
As extensively discussed by SWM, the integrated intensity obtained
from our model AGN luminosity function is also consistent with all the available
data on the cosmic X-ray background for energies above 1 keV.
Our qualitative conclusions would not change if we adopted
the Hopkins et al. (2007) luminosity function in place of SWM,
though best-fit parameter values would be somewhat different.
\section{Effects of a broad $P(\lambda)$}\label{sec|broadPL}
In this section and the ones that follow, we will compare predictions
of our models to a variety of observational constraints. If the observational
uncertainties were described by well defined statistical errors, then
the natural approach would be to determine best-fit model parameters
via $\chi^2$ minimization and compare different models via
$\Delta\chi^2$. We have taken this approach in some of our previous
studies when we were focusing on a specific observational constraint
and a limited class of models. However, the uncertainties in the
observational constraints we consider --- AGN luminosity functions,
the local black hole mass function, distributions of Eddington ratios,
and duty cycles estimated from active galaxy fractions --- are in all
cases dominated by systematic errors, and in some cases even the
rough magnitude of these systematics is difficult to estimate.
We discuss these issues in the text and Appendices below, and the papers cited
for each observable often discuss systematic uncertainties at length.
Given this situation, we adopt a philosophy of ``qualitative comparison
to quantitative data''. We infer model parameters based on an overall
match to one or more sets of observational constraints, but we do not
attempt formal $\chi^2$-minimization because the errors themselves
are not well defined enough to do so. We note where models fit observations
within a plausible range of systematic uncertainties, and where
they do not. Our objective is to delineate the global characteristics
that successful black hole accretion models must possess to reproduce
the range of observables now emerging from deep, large area surveys of
the AGN and galaxy populations.
\subsection{Input Eddington ratio distributions}\label{subsec|PLmodels}
\begin{figure*}
\includegraphics[width=15truecm]{Figure2.eps}
\caption{\emph{Left panels}: Predicted luminosity function at $z=2$
(top) and at $z=1$ (bottom)
for the three $P(\lambda)$\ distributions of Fig.~\ref{fig|comparePLambda},
as labelled,
when the {\it same}
black hole mass function $\Phi_{\rm BH}(M_{\rm BH},z)$\ and duty cycle $U(M_{\rm BH},z)$\ are used, in this
case the
one predicted by the $\delta$-model.
Points show the data compilation of SWM.
\emph{Right panels}: predicted duty cycle $U(M_{\rm BH},z)$\ at the same redshifts
$z=2$ and $z=1$
when, instead, the
AGN luminosity function is \emph{fixed} to the one in SWM,
as in our self-consistent models.
Allowing a large fraction of sub-Eddington accretion rates,
as in the G+P model, increases the probability and thus the duty cycle
for more massive black holes to be active
at different luminosities.}
\label{fig|predictedLFandUmbh}
\end{figure*}
\begin{figure*}
\includegraphics[width=17truecm]{Figure3.eps}
\caption{The \emph{left}, \emph{middle}, and \emph{right} columns refer
to outputs
of the $\delta$, G, and G+P models, respectively, whose $P(\lambda)$\ distributions
are shown in Fig.~\ref{fig|comparePLambda}.
\emph{Upper panels}: Predicted black hole mass function at $z\sim 0$
(\emph{solid} lines) compared to the local estimates by SWM (\emph{grey
area}),
and Vika et al. (2009; \emph{squares} with error bars,
with the \emph{open} symbols indicating the estimates derived from
galaxies below their reliability limit). Other lines
show the predicted black hole mass function at earlier redshifts, as
labelled. \emph{Bottom panels}: Corresponding duty cycles (for all
sources accreting above any $\lambda>0$)
as a function of black hole mass at different redshifts, as labelled. For
all models a radiative efficiency of $\epsilon=0.06$ has been assumed.}
\label{fig|BHMFandDuty}
\end{figure*}
In SWM we discussed accretion models for black holes with a single value
of $\lambda$ , examining the
effects of redshift and mass dependence of this value of $\lambda$\
and describing the impact of uncertainties
in the input AGN luminosity function related to its observational
determination, obscured fractions, and bolometric corrections.
From the match to the local black hole mass function, we were able to set
constraints on the average input radiative efficiency and evolution in the
characteristic $\lambda$.
In this section we study the impact of broadening the input
$P(\lambda)$\ distribution. To this purpose we adopt and compare the outputs
from three different distributions (shown in Fig.~\ref{fig|comparePLambda}):
\begin{itemize}
\item the \textbf{$\delta$-model}, centered on $\log \lambda_c$$=-0.6$ (dot-dashed line
in panel \emph{a}; we choose this particular value because it provides
a good match to the local black hole mass function);
\item the \textbf{G-model}, a Gaussian in $\log\lambda$
with dispersion of $\Sigma_{\lambda}$$=0.3$
dex and centered around the same value of $\log \lambda_c$$=-0.6$ (long-dashed line
in panel \emph{a});
\item the \textbf{G+P-model}, characterized by the same Gaussian plus a
power-law with slope $\alpha=-0.3$ normalized
to have the same value as the Gaussian at $\log \lambda=\log \lambda_c-
\log \Sigma_{\lambda}$ (panel \emph{b}).
\end{itemize}
The G+P distribution has a shape close to the one
inferred by Kauffmann \& Heckman (2009; see also Aird et al. 2011) in the local Universe from
the Sloan Digital Sky Survey (SDSS),
and it is also consistent with some theoretical and semi-empirical
expectations (e.g., Merloni \& Heinz 2008; Yu \& Lu 2008; Hopkins \&
Hernquist 2009; Shen 2009).
In particular, the power-law component could represent the effect of either
a steady decline in the accretion rate after a near-Eddington growth phase or
a second mode of AGN fueling triggered by secular instabilities instead
of major mergers.
As discussed by SWM and demonstrated further below, the chosen value of $\log \lambda_c$\ produces reasonable agreement between the local and accreted mass function, especially around $(1-3)\times 10^8\, $$M_{\odot}$\ where the former is best determined. Our chosen value of $\alpha$, on the other hand, was mainly derived a posteriori after some trial and error. Higher values of $\alpha$ can induce unphysical duty cycles $U(M_{\rm BH},z)$\ $>1$
in some mass bins during the evolution,
while lower values give results that are not much different from the G-model alone.
For all models we adopt a radiative efficiency of $\epsilon=0.06$, unless otherwise stated, within the range of values favoured by SWM.
In future sections and figures we will consider many variants on
these three basic models.
For reference, we provide a full list of models and their basic
properties in Table~\ref{table|models}. We summarize the comparison
between our six primary models --- the three introduced here, plus versions
that introduce redshift and mass dependence of $P(\lambda)$\ ---
and observational data in Table~\ref{table|models2} (Section~\ref{sec|discu}).
\begin{table*}
\begin{tabular}{|l|l|r|}
\hline
Reference Models & shape \& properties & Section/Equations \\
\hline
\hline
$\delta$ & $\delta$-function & Sec.~\ref{subsec|PLmodels} \\
G & Gaussian & Sec.~\ref{subsec|PLmodels} \\
G+P & Gaussian+Power-Law & Sec.~\ref{subsec|PLmodels} \\
G(z) & Gaussian+$\lambda_c(z)$ & Eq.~\ref{eq|PLz} \\
G(z,$M_{\rm BH}$) & Gaussian+$\lambda_c(z,M_{\rm BH})$+low-$z$
broadening+mass-dependent $\epsilon$ & Eqs.~\ref{eq|PLz}, \ref{eq|PLmassDependent}, \ref{eq|Sigma}, \ref{eq|radefficiencyMass} \\
G+P(z,$M_{\rm BH}$) & Gaussian+$\lambda_c(z,M_{\rm BH})$+$z$-dependent
Power-Law+mass-dependent $\epsilon$ & Eqs.~\ref{eq|PLz}, \ref{eq|PLmassDependent}, \ref{eq|LambdaMin}, \ref{eq|radefficiencyMass}\\
\hline
\hline
Test Models & shape \& properties & Ref. Eqs. \\
\hline
\hline
G(z,$M_{\rm BH}$)+constant $\epsilon$ & & \\
G(z,$M_{\rm BH}$)+$\epsilon(z)$ & G(z,$M_{\rm BH}$)+z-dependent $\epsilon$ & Eq.~\ref{eq|radEfficiencyReds}\\
G(z,$M_{\rm BH}$)+$K(\lambda)$ &
G(z,$M_{\rm BH}$)+$\lambda$-dependent bolometric correction
& Eq.~\ref{eq|BClambda} \\
G(z,$M_{\rm BH}$)+P($z<0.7$) & G(z,$M_{\rm BH}$)+Power-Law only at $z<0.7$ & \\
\hline
\end{tabular}
\caption{List of models explored in this work along with basic explanation and Sections and Equations in which they are first introduced. The models are divided into
two groups, the ``Reference Models'', i.e., the three ones introduced and discussed in Section~\ref{subsec|PLmodels}, and the ``Test Models'' which are simply variants of the Reference Models and progressively introduced in the rest of the paper.
The Gaussian always has
dispersion of $\log \Sigma_{\lambda}=0.3$ dex except for the models with
broadening, for which
$\log \Sigma_{\lambda}$ steadily increase with decreasing redshift as
in Eq.~(\ref{eq|Sigma}).
The redshift and mass
dependence in the characteristic Eddington ratio $\lambda_c$ are defined
in Eqs.~(\ref{eq|PLz}) and (\ref{eq|PLmassDependent}). If no evolution
is assumed then $\log \lambda_c=-0.6$. The power-law component of the
Eddington distribution is always normalized
to have the same value as the Gaussian at $\log \lambda=\log \lambda_c-
\log \Sigma_{\lambda}$, and a slope in $\log \lambda$ of $\alpha=-0.3$.
In the last quoted model the power-law has slope $\alpha=-0.9$ and is normalized
to have the same value as the Gaussian at $\log \lambda=\log \lambda_c-
0.2\log \Sigma_{\lambda}$.
The input radiative efficiency can be constant or vary with mass or redshift
(see Eqs.~\ref{eq|radefficiencyMass} and~\ref{eq|radEfficiencyReds})
as detailed for each model.}
\label{table|models}
\end{table*}
The left panels of Fig.~\ref{fig|predictedLFandUmbh} illustrate the effect
of varying the input Eddington ratio distribution $P(\lambda)$\ at fixed duty
cycle $U(M_{\rm BH},z)$\
and fixed black hole mass function $\Phi_{\rm BH}(M_{\rm BH},z)$, at $z=2$ and $z=1$.
The luminosity function
in each case is computed via the convolution of the respective $P(\lambda)$\
distributions
with the product of the duty cycle and black hole mass function
(Eq.~\ref{eq|PhiLLambda}). In the
case of the $\delta$-model, the match to the input AGN luminosity
function (Section~\ref{subsec|AGNLF}) is exact.
For these panels we \emph{fix} $U(M_{\rm BH},z)$\ and $\Phi_{\rm BH}(M_{\rm BH},z)$\ to those predicted by the $\delta$-model
(evolved in accord with the observed luminosity function),
so that we can isolate the impact of varying the input $P(\lambda)$\ distribution.
Switching from a $\delta$-function $P(\lambda)$\ to a Gaussian of
$\Sigma_\lambda=0.3\,$dex has a
mild impact on the predicted luminosity function.
However, adopting the G+P distribution at fixed $U(M_{\rm BH},z)$\
drastically lowers the luminosity function at high $L$, since the
probability is dominated by low Eddington ratios. The shape of
$\Phi(L,z)$ parallels that of the other models above the break
luminosity, but the G+P model has a steeper faint-end slope
reflecting the contribution of low-$\lambda$ activity.
The procedure we follow in the rest of this paper is to compute $U(M_{\rm BH},z)$\
for a given input $P(\lambda)$\ distribution in a way to reproduce the observed AGN
luminosity function.
The right panels of Fig.~\ref{fig|predictedLFandUmbh} show, at the same
redshifts $z=2$ and $z=1$, the duty cycles inferred by matching the SWM
luminosity function when adopting the three $P(\lambda)$\ distributions
and the same underlying $\Phi_{\rm BH}(M_{\rm BH},z)$, again that of the $\delta$-model.
The duty cycles $U(M_{\rm BH},z)$\ for the $\delta$\ and G models are similar,
but they are several times larger for the G+P model
at high masses.
As described in SWM, the duty cycles must decrease with mass in order to
reproduce the ``downsizing'' effect in AGN evolution (the characteristic
AGN luminosity increases with redshift). In the G+P model the duty
cycle can reach values close or around unity at $z \lesssim 2$ and masses $\sim 10^7 M_{\odot}$, although with low values of $\lambda$ for the majority of them. The decline of the duty
cycle in this model at masses lower than
$M_{\rm BH} \la L_*/(l\lambda_c)$, where $L_*$ is the break of the luminosity
function (Eq.~[\ref{eq|U03}]), is induced by the presence of
many AGNs radiating at low $\lambda$ from high mass black holes, which can
already account for the low-luminosity AGNs.
However, the detailed form of this decline is sensitive to the precise
shape of the AGN luminosity function and our assumed $P(\lambda)$.
\begin{figure*}
\includegraphics[width=17truecm]{Figure4.eps}
\caption{\emph{Upper panels}: Growth rate of the integrated black hole
mass density as a function of time (\emph{solid} lines) and relative
contributions of black holes of different final mass, as labelled;
the \emph{grey area} marks the 3$\sigma$ uncertainty region of the
cosmological star formation rate as inferred by Hopkins \& Beacom (2006),
scaled by a factor of $6.5\times 10^{-4}$. \emph{Lower panels}: predicted
cumulative black hole mass density as a function of redshift (\emph{solid}
lines), and contributions of black holes of different mass, as labelled.
The \emph{grey bar} indicates the values and systematic
uncertainties in the total local mass density in black holes estimated
by SWM}.
\label{fig|RhoBHz}
\end{figure*}
\subsection{Global Accretion Histories}\label{subsec|AccretionHistories}
Fig.~\ref{fig|BHMFandDuty} shows the
evolution of the black hole mass functions and duty cycles
for the $\delta$, G, and G+P models, respectively.
In the upper panels, grey bands show the estimate of the local black
hole mass function by SWM. The width of this band already encompasses
a number of systematic uncertainties, but, as SWM discuss, the
inferred mass function for $M_{\rm BH} < 10^8 M_\odot$ is sensitive
to uncertainties in the treatment of spiral galaxy bulges.
Vika et al.\ (2009) have tried to address this problem by
estimating the local black hole mass function on an object by object
basis, i.e., by computing the
bulge fraction for each galaxy in a large sample, assigning
black hole masses from an $M_{\rm BH}$-$L$ relation, and then computing
the black hole mass function applying the $V/V_{\rm max}$ method.
Their result, shown with open and filled squares in
Fig.~\ref{fig|BHMFandDuty}, agrees well with SWM at high masses,
while it turns over at $M < 10^8 M_\odot$ rather than continuing to rise
to lower masses.
We thus consider Vika et al.'s and SWM's results to broadly bracket
the still remaining uncertainties in the determination
of the local black hole mass function, and we will use both as a reference
when comparing with models.
Moving from the $\delta$-model to the G-model makes almost no difference
to the evolution of the black hole mass function or to its $z=0$
value. Both predictions are insensitive to our assumed initial
conditions at $z=6$ because even by $z=4$ the accumulated mass from
accretion greatly exceeds that in the initial seed population.
The G+P model differs in this regard because reproducing the $z=6$
luminosity function even with a duty cycle of 0.1 requires a high
space density of seed black holes, since many black holes are
active at low $\lambda$ values that do not contribute to the observed
range of the luminosity function. Once the evolved black hole mass
function substantially exceeds the seed population, evolution is
similar to that of the $\delta$\ and G models.
The clustering of quasars is a diagnostic for the space density
of the underlying black hole population, as more numerous black holes
must reside in lower mass halos that are less strongly clustered
(Haiman \& Hui 2001; Martini \& Weinberg 2001). Applying this idea
to the strong clustering measured for $z\approx 0.4$ quasars by
Shen et al.\ (2007) in the SDSS, White et al.\ (2008) and
Shankar et al.\ (2010b) find that duty cycles close to unity
are required assuming that these highly luminous quasars have
$\lambda \ga 0.1$. This finding is at odds with the high black hole density
required for the G+P model at high redshift. To quantify this
statement, we have computed the large scale bias of the G+P model
using a formalism that we will describe more fully in a future paper.
In brief, we match black holes to halos and subhalos
via cumulative number matching
(similar to Conroy \& White 2012), including 0.3-dex scatter in black
hole mass about the mean relation
(see Appendix~\ref{Appendix|MbhMhaloRelation}), then assign each
black hole a luminosity based on the duty cycle and the Eddington
ratio distribution $P(\lambda|M_{\rm BH},z)$. We can then compute the mean large
scale halo bias for black holes in a luminosity range. For the G+P
model, which has high black hole space density and large scatter
between luminosity and halo mass, we predict a bias of 7.5 for quasars
with $\log L > 47$ at $z=4$, which is more than $2\sigma$ discrepant
with the reference value of $12.96 \pm 2.09$ given by
Shen et al.\ (2009), though it is marginally consistent with
their lowest estimate.
We conclude that broad $P(\lambda)$\ distributions at very high redshifts
are observationally disfavoured, a conclusion in qualitative
agreement with the findings of Cao (2010). If $P(\lambda)$\ does have a
long tail to low $\lambda$ values, it must develop after the
earliest stages of black hole growth;
we will consider models with growing power-law tails
in Section~\ref{sec|ActiveFractions}.
Duty cycles $U(M_{\rm BH},z)$\ tend to decline with redshift and with mass, for the
reasons already explained in Fig.~2. We note again the higher values
of the duty cycle in the G+P model at all redshifts. The high-redshift
curves are still affected by the initial conditions in the black hole
population at low masses.
\begin{figure*}
\includegraphics[width=12truecm]{Figure5.eps}
\caption{Eddington ratio distributions predicted by the G
(\emph{solid} lines) and G+P (\emph{long-dashed} lines) models, compared
to the K06 data (\emph{grey histograms}) for black
holes in the range $10^7<M_{\rm BH}/M_{\odot}<10^{10}$ at different redshift
and luminosity bins, as labelled. The predictions have been estimated
at the average redshifts of $z=1$ and $z=2$ for the low and high-$z$
subsamples, respectively, although the predictions do not strongly depend
on these choices if no redshift dependence in the $P(\lambda)$\ is included. The
\emph{cyan long-dashed line} in the lower-right panel is the G+P model
prediction when no lower luminosity threshold is considered.}
\label{fig|CompareKollmeier1}
\end{figure*}
Fig.~\ref{fig|RhoBHz} tracks the overall growth of the black hole
population in bins of mass. In the upper panels, solid lines
show the growth rate of the integrated
black hole mass density $d\rho_{\rm BH}/dt(z)$\
as a function of redshift, while the other lines show the mass
density accreted in selected bins
of \emph{current} black hole mass, as labelled. As already noted by several
groups (e.g., Marconi et al. 2004; Merloni et al. 2004; Merloni \& Heinz
2008; SWM),
the total $\rho_{\rm BH}(z)$\ closely matches the shape of the cosmological star formation
rate (SFR), here taken from Hopkins \& Beacom
(2006; with the gray area marking their 3-$\sigma$ contours)
and re-scaled by an \emph{ad hoc} factor of $6.5\times 10^{-4}$, close
to the ratio between black hole mass and stellar mass measured in the local
Universe (e.g., H\"{a}ring \& Rix 2004).
All three models again show a clear signature of downsizing in their
accretion histories.
The accretion onto the very massive
black holes with $\log M_{\rm BH}/M_{\odot}>8$ always peaks at $z\sim 2$,
concurrent with the peak in the emissivity of luminous optical quasars
(e.g., Osmer 1982; Richards et al. 2006; Croom et al. 2009). The less
massive black
holes with $7<\log M_{\rm BH}/M_{\odot} < 8$ are characterized by a much
broader peak centered at $z\sim 1-1.5$.
The lower panels show the cumulative mass density
of all black holes with mass $10^6<M_{\rm BH}/M_{\odot}<10^{10}$ (solid lines),
and the mass density accreted onto black holes of different current mass,
as labelled.
By construction, all models share the same radiative efficiency and therefore
accumulate the same total mass densities at any time. They
differ only slightly
with respect to the total mass accumulated in different mass
bins. The solid grey square indicates the systematic uncertainties
in the total local mass density in black holes estimated by SWM.
Figures~\ref{fig|BHMFandDuty} and~\ref{fig|RhoBHz} show that the
evolution of the black hole mass function is insensitive to the
shape of $P(\lambda)$, provided the characteristic value $\lambda_c$ and
the input AGN luminosity function are held fixed.
This reassuring result
indicates that earlier studies assuming a single $\lambda$ value
(e.g., Marconi et al.\ 2004; Shankar et al.\ 2004; SWM) reached
robust results for mass function evolution.
This is essentially because in our G+P model,
with power-law slope $\alpha=-0.3$,
the emissivity per logarithmic
interval of $\lambda$ is dominated by the Gaussian peak
(Fig.~\ref{fig|comparePLambda}b).
A steeper slope would yield duty cycles greater than unity and is
therefore not allowed.
Even our G+P model is disfavored at high redshifts because
the observed quasar clustering implies that the most massive black
holes should be active at large values of $\lambda$ with high duty
cycles. Our model results are obviously dependent on
the value of $\lambda_c$, which changes the location of the break
in the black hole mass function, and the value of $\epsilon$, which
affects the normalization
(see SWM and Section~\ref{sec|ActiveFractions} below).
\begin{figure*}
\includegraphics[width=16truecm]{Figure6.eps}
\caption{Eddington ratio $P(\lambda)$\ distributions predicted by the
G (\emph{solid} lines) and G+P (\emph{long-dashed} lines)
models at lower redshifts. \emph{Left panel}: Predictions
compared to the H09 data (\emph{grey histogram})
for black holes in the range $10^7<M_{\rm BH}/M_{\odot}<10^{10}$
with $\log L > 43.5$ at $z=0.5$.
\emph{Right panel}: Predictions compared to the
KH09 data (\emph{grey histogram})
for black holes in the range $10^7<M_{\rm BH}/M_{\odot}<10^8$ with
$\log L> 42.5$ at $z=0.2$.}
\label{fig|CompareKauff1}
\end{figure*}
\subsection{Comparison with measured $P(\lambda)$\
distributions}\label{subsec|CompareWithObservedPL}
Drawing on virial mass estimators grounded in reverberation mapping studies
(e.g., Wandel, Peterson \& Malkan 1999; Vestergaard 2002; McLure \&
Jarvis 2004)
and the availability of large samples of quasar spectra
from wide field surveys, several groups have inferred the distribution
of Eddington ratios in different ranges of redshift and luminosity.
Kollmeier et al. (2006; hereafter K06) measured $P(\lambda)$\ from the
AGN and Galaxy Evolution Survey (AGES; Kochanek et al. 2011), finding that
luminous AGNs at $0.5<z<3.5$ have a quite narrow range of Eddington
ratios, with a peak at $\lambda \sim 0.25$ and a dispersion
(including observational errors) of $\sim 0.3$ dex.
Netzer et al. (2007), analyzing quasar samples
from the SDSS,
found a similar result, with a slightly larger dispersion, from a sample
centred at $z\sim 2.5$. Netzer \& Trakhtenbrot (2007, see also
Vestergaard 2004; McLure \& Dunlop 2004), analyzing nearly ten
thousand SDSS quasars in the redshift range $0.3\lesssim z
\lesssim 0.75$, confirmed the log-normal shape of the Eddington ratio
distribution and also found evidence for a
significant decrease with time in the characteristic $\lambda_c$
defining the peak of the distribution.
K06 also found a factor $\sim 3$ decrease in
the peak value of $\lambda$ between high and low redshifts for
the more massive black holes in their sample.
Many other studies of the Eddington ratio distributions of quasars
and active galaxies have found
evidence for time evolution of $\lambda_c$,
and in some cases for mass-dependence
(e.g., McLure \& Dunlop 2004; Vestergaard 2004; Heckman et al. 2005;
Ballo et al. 2007; Babi\'{c} et al. 2007; Bundy et al. 2008;
Rovilos \& Georgantopoulos 2007; Cao \& Li 2008;
Fine et al. 2008; Shen et al. 2008; Greene et al. 2009; Hickox et al. 2009;
Kauffmann \& Heckman 2009; Kelly et al. 2010; Shankar et al. 2010e;
Steinhardt \& Elvis 2010a,b; Trakhtenbrot et al. 2011; Willott et al. 2010a,b; Aird et al. 2011; Shen \& Kelly 2011).
However, sample selection, observational noise, and intrinsic
scatter in black hole mass estimators can all have strong effects
on the apparent distribution of Eddington ratios
(e.g., Lamastra et al. 2006; Lauer et al. 2007; Marconi et al. 2008;
Shen et al. 2008; Netzer 2009; Shen \& Kelly 2010; Rafiee
\& Hall 2011).
At low redshifts, where samples can probe to much lower AGN luminosities,
it is clear that the distribution of Eddington ratios is much broader
than the roughly log-normal distribution found for optically luminous
quasars at high redshift.
As a set of representative examples of the literature,
we consider the Kollmeier et al. (2006; K06 hereafter) results for broad-line
optically luminous quasars at $z > 0.5$, the Hickox et al. (2009; H09
hereafter) results
at $z\approx 0.5$ for a sample selected by X-ray, infrared,
and radio emission, and Kauffmann \& Heckman's (2009; KH09 hereafter)
analysis
of the local AGN population in the SDSS, with an effective
redshift $z \approx 0.2$.
To make a close comparison of our models to
observed samples of AGNs,
we compute Eddington ratio distributions conditioned
to active AGNs with luminosities between $L_{\rm min}$ and $L_{\rm max}$, $P_{\rm obs}$, as
\begin{eqnarray}
P_{\rm obs}(\lambda|z,L_{\rm min},L_{\rm max})=\\ \nonumber
A\sum_k \sum_j
P(\lambda=L_k/(M_{\rm BH,j}l)|z,M_{\rm BH,j}) N_{\rm act}(M_{\rm BH,j},z)~,
\label{eq|ComputePLzM}
\end{eqnarray}
where $N_{\rm act}(M_{\rm BH},z)$\ is defined in Eq.~\ref{eq|NactDef}.
The sums in the above equation are extended to all bins of
black hole mass, and to all bins of luminosity
in the assumed observed range, $L_{\rm min}<L_k<L_{\rm max}$.
Since we are interested here in the {\it shapes} of the predicted Eddington distributions,
the constants $A$ are chosen to renormalize the
$P_{\rm obs}(\lambda|z,L_{\rm min},L_{\rm max})$\ to match the peaks of the observed distributions.
Fig.~\ref{fig|CompareKollmeier1} compares
the Eddington ratio distributions predicted by the G
and G+P models to the K06 data,
for black holes in the range $10^7<M_{\rm BH}/M_{\odot}<10^{10}$
in four different bins of redshift and luminosity.
The predictions have been computed at the redshifts
of $z=1$ and $z=2$ for comparison to the $z<1.2$ and $z>1.2$ subsamples,
respectively. The predictions do not
strongly depend on these choices, since
$P(\lambda)$\ has no explicit redshift dependence.
Both models predict, essentially by construction, a roughly
log-normal $P(\lambda)$\ with a typical $\lambda_c$ that agrees with the
K06 measurements. Most importantly, the
power-law tail of the G+P model does {\it not} lead to
substantial disagreement
because the fraction of quasars of a fixed luminosity with low values of $\lambda$ is much less than the fraction of black holes of a fixed mass with the same low values of $\lambda$.
In fact, the $\lambda$ distribution for all quasars with any
luminosity (cyan dashed curve
in the lower right panel) has the same shape as our assumed intrinsic distribution in Figure 1a.
The most significant disagreement with the data is the
high characteristic $\lambda$ for $45.5 < \log L < 46$ at $z<1.2$
(lower left panel), a consequence of keeping the model $\lambda_c$
fixed instead of allowing redshift evolution.
Fig.~\ref{fig|CompareKauff1} compares the predicted and observed
$\lambda$ distributions for
the lower redshift samples of H09 at
$z\approx 0.5$ and KH09 at $z\approx 0.2$.
To roughly take into account the effective limits in black hole mass
and luminosity of the observational samples,
for the former we consider all black holes with $M_{\rm BH} > 10^7\, M_{\odot}$
shining at $L > 10^{43}{\rm erg \, s^{-1}}$, while in the latter case we
adopt the restricted range $10^7 < M_{\rm BH}/M_{\odot} < 10^8$
shining above $L=10^{42.5}\, {\rm erg \, s^{-1}}$.
The match to the observations is poor in both cases.
The distributions predicted by the G model are much narrower than
the observed distributions, and they peak at a value
$\lambda_c \approx 0.25$ that is roughly ten times
higher than the high-$\lambda$ peaks of the observed
distributions.
The power-law tail in the G+P model is in rough agreement
with the observed distributions at low $\lambda$,
but the peak of the Gaussian component remains discrepant.
Fitting these observations, even approximately, requires evolution
of $\lambda_c$, decreasing towards low redshift.
We return to this point in Section~\ref{subsec|EddingtonRatioDistributions}
below, after first finding independent evidence for evolution
in $\lambda_c$ from observed active galaxy fractions.
\begin{figure*}
\includegraphics[width=17truecm]{Figure7.eps}
\caption{
Comparison of predicted and observationally estimated duty
cycles at low redshift.
\emph{Left panel}: Data points from Goulding et al.\ (2010)
compared to model predictions at $z=0.2$ as labelled,
with a luminosity threshold $\log L >41.5$.
\emph{Middle panel}: Data points from Kauffmann et al. (2003)
compared to model predictions, with a luminosity threshold
$\log L > 43.5$.
\emph{Right panel}: Data points from Schulze \& Wisotzki
(2009) compared to model predictions, with an {\it Eddington ratio} (not luminosity) threshold $\lambda>0.01$
and a correction factor for the Type 1 Broad Line AGNs taken from Greene \& Ho (2009). The three
lowest mass points include the Schulze \& Wisotzki (2010)
incompleteness corrections, which are negligible at
higher mass.
}
\label{fig|DutyCyclez0}
\end{figure*}
\section{Active Galaxy Fractions: redshift and mass-dependence of
$P(\lambda)$}\label{sec|ActiveFractions}
Duty cycles are a basic prediction of continuity equation
models as discussed in Section~\ref{sec|formalism}.
If we assume that most or all massive galaxies contain black holes,
as supported by observations of the local universe (see, e.g., Ferrarese \& Ford 2005, and references therein), then
active galaxy fractions can be used as an observable proxy for
black hole duty cycles.
However, active galaxy fractions depend on the
luminosity or Eddington ratio threshold of each specific
observational sample:
lower thresholds obviously increase the
active fraction. Therefore, in order to properly compare model predictions
to available data sets, we
define the fraction of black holes shining above a luminosity
threshold $L_{\rm min}$, or observed duty cycle $U_{\rm obs}$, as
\begin{eqnarray}
U_{\rm obs}(M_{\rm BH},L_{\rm min},z)=\int_{\log \lambda_{\rm min}}^{\infty}
P(\lambda|M_{\rm BH},z) U(M_{\rm BH},z)d \log \lambda\, , \\
\lambda_{\rm min} = L_{\rm min}/(l M_{\rm BH})\, . \nonumber
\label{eq|DutyCycleAboveL}
\end{eqnarray}
In this Section and Section~\ref{sec|OtherConstraints} we will
introduce versions of our reference models that incorporate redshift
and mass dependence of $P(\lambda)$, and we will consider a variety
of ``test'' models that illustrate specific points. The key features
of these models are summarized in Table~\ref{table|models}, and we
will summarize the qualitative successes and failures of the
reference models in Table~\ref{table|models2} below (Section~\ref{sec|Discussion}).
These model changes generally have little impact on the overall mass
accretion histories, so we will not repeat the analysis shown previously
in Figures~\ref{fig|BHMFandDuty} and~\ref{fig|RhoBHz} but instead focus
on predictions where the new models differ significantly from those
described in Section~\ref{sec|broadPL}. Despite systematic uncertainties
in the observational constraints, even qualitative trends with
redshift and black hole mass are enough to provide strong model tests.
\subsection{Observational Estimates}\label{subsec|ObservationalEstimates}
We first summarize the observational estimates of AGN fractions
that we adopt for our model comparisons, with further details of
some of these estimates given in
Appendix~\ref{Appendix|AGNdutyCycles}.
For high luminosity, broad-line (Type 1) AGN in the local Universe,
Schulze \& Wisotzki (2010) have estimated the mass function of
active black holes at $z< 0.3$ by applying linewidth mass estimators
to quasars identified in the Hamburg/ESO objective prism survey.
They divide their active black hole density by the total black
hole space density in the Marconi et al.\ (2004) mass function to
derive an active fraction. Schulze \& Wisotzki impose a threshold
of $\lambda \geq 0.01$ in their estimated Eddington ratio to define
active systems. Host galaxy contamination causes incompleteness
in their AGN catalog below $M_{\rm BH} \approx 10^{7.5}M_\odot$
(for $\lambda \geq 0.01$). In the right panel of
Fig.~\ref{fig|DutyCyclez0} we plot the points from their
Fig.~12, correcting their three lowest $M_{\rm BH}$ points for
incompleteness as suggested by the authors.
According to their simulations, higher $M_{\rm BH}$ points should be unaffected by incompleteness,
and we have omitted points below $M_{\rm BH} = 10^7 M_\odot$ where
incompleteness corrections become large.
The active black hole fractions estimated by Schulze \& Wisotzki (2010)
are low, only $\sim 2 \times 10^{-4}$ at $M_{\rm BH} = 10^9 M_\odot$,
rising to $\sim 10^{-3}$ at $M_{\rm BH} = 10^8 M_\odot$.
\begin{figure*}
\includegraphics[width=17truecm]{Figure8.eps}
\caption{Predicted and observationally estimated duty cycles
at redshifts $z=0.5$, $z=1.3$, and $z=2.3$.
In all panels, \emph{grey bands} show the estimates of
Xue et al. (2010).
Points in the \emph{middle} panel, points are from Bundy et al.\ (2008).
Points in the \emph{right} panel come from
Erb et al. (2006; \emph{filled small squares}), Kriek
et al. (2007; \emph{filled triangles}), Alexander et al. (2005; 2008;
\emph{filled circle}), and Caputi et al. (2006; \emph{semi-filled circle}).
}
\label{fig|DutyCycleHighz}
\end{figure*}
Kauffmann et al.\ (2003) estimate much higher active fractions
for narrow line (Type 2) AGN identified by applying emission
line diagnostics to SDSS spectra in the SDSS main galaxy redshift
survey. Since Kauffmann et al.\ (2003) begin with a galaxy catalog,
quasars that are bright enough to appear as point sources (rather than
extended sources) in SDSS imaging are excluded.
They estimate their completeness limit at $10^7\, L_\odot$ in
[OIII] luminosity, which we convert to an approximate
$L_{\rm min} = 10^{43.5}{\rm erg \, s^{-1}}$ by adopting an extinction-corrected
bolometric correction of $\sim 800$ (both Kauffmann et al.\ 2003 and KH09
adopt extinction-corrected luminosities).
Points in the middle panel of Fig.~\ref{fig|DutyCyclez0}
are taken from Fig.~5 of Kauffmann et al.\ (2003),
where we have simply converted stellar mass to black hole
mass by multiplying by $1.6\times 10^{-3}$ (Magorrian et al.\ 1998).
This crude conversion should be reasonably accurate above
$M_* \sim 10^{11.5} M_\odot$, but it probably overestimates
the black hole mass for lower mass galaxies that are no longer
bulge-dominated. The Kauffmann et al. (2003) active fractions
rise from $\sim 2\%$ at $M_{\rm BH} = 10^9 M_\odot$ to
$\sim 10\%$ at $M_{\rm BH} = 10^8 M_\odot$, then decline towards
lower $M_{\rm BH}$. The $10^{43.5}{\rm erg \, s^{-1}}$ luminosity threshold
corresponds to $\lambda = 0.01$ at $M_{\rm BH} \approx 10^{7.5} M_\odot$,
so at higher masses the Kauffmann et al.\ (2003) estimates correspond
to a lower $\lambda$-threshold than Schulze \& Wisotzki's.
The $1.5$ to 2 order-of-magnitude gap between the Kauffmann et al.\ (2003)
and Schulze \& Wisotzki (2009) active fractions
can be partly explained by the different abundances of
broad line and narrow line AGN at low redshift. The abundance ratio of
the two types of AGN has been addressed by
Greene \& Ho (2009; a correction of Greene \& Ho 2007), who
analyze all SDSS spectroscopic objects (targeted as galaxies
or quasars) in Data Release 4 (Adelman-McCarthy et al.\ 2006)
to define a sample of 8,400 broad-line AGN at $z \la 0.3$.
Their estimated active fractions (in their Fig.~11) are
close to those of Schulze \& Wisotzki. They find a gap
of 0.9-1.3 dex between their active black hole mass function for
broad-line AGN and the corresponding result (Heckman et al. 2004)
for SDSS narrow-line AGN (see Greene \& Ho 2009, Figure 10),
which implies a correction of $\sim 10-20$ for the ratio of
Type I to Type II AGN at low redshift.
The remaining factor of $5-10$ may be accounted for by
the effects of scatter in bolometric corrections,
different luminosity thresholds, and systematic offsets in the
black hole mass function estimates.
Note that scatter in $L_{\rm bol}/L_{\rm [OIII]}$ could
make the effective luminosity threshold of the Kauffmann et al. (2003)
data lower than we have assumed, in particular if the luminosity function
is affected by a steeply rising low-$\lambda$ tail in $P(\lambda)$.
For studies that probe to lower AGN luminosities, active galaxy
fractions are even higher than those of Kauffmann et al. (2003).
For example, Ho (2004) argues that at least 40\% of local
bulge-dominated galaxies host a low luminosity AGN and/or a LINER,
and Grier et al. (2011) find nuclear X-ray sources in $\sim 60\%$
of galaxies in the SINGS survey (Kennicutt et al. 2003).
As a representative example of low luminosity statistical studies,
we show in the left panel of Fig.~\ref{fig|DutyCyclez0} the data
of Goulding et al. (2010, their Fig.~5), based on an X-ray and IR
census of a volume-limited sample of galaxies with $D < 15$ Mpc.
Their active fractions range from 14\% to 37\% over the black
hole mass range $10^6 - 10^{8.5}\, M_{\odot}$, albeit with large
statistical errors reflecting the small sample size.
The luminosity threshold of this sample is very low,
about $L_{\rm min} = 10^{41.5}\, {\rm erg \, s^{-1}}$ in bolometric units (their
Fig.~3).
At higher redshifts we take as our primary data set the measurements
reported by Xue et al.\ (2010) based on moderate luminosity X-ray
AGN in the Chandra Deep Fields. We reproduce results from their
Fig.~14 (the upper, orange bands) as grey bands in the three
panels of Fig.~\ref{fig|DutyCycleHighz}, converting their
X-ray luminosity thresholds to bolometric luminosity thresholds
with our luminosity-dependent bolometric correction (Section~\ref{subsec|AGNLF}).
We again convert stellar
masses to black hole masses by simply multiplying by the
Magorrian et al. (1998) factor of $1.6\times 10^{-3}$. This scaling should
be taken with a grain of salt, as we are using total stellar masses
rather than bulge masses and ignoring possible redshift evolution
of the scaling factor. However, these uncertainties should not
affect the key lessons that we take from Xue et al.\ (2010): at all
three redshifts, the duty cycle for BHs shining above $\log L>43.3$
increases with increasing $M_{\rm BH}$
(indicated in the data by increasing $M_*$), and for the most massive
galaxies active fractions are $10-50\%$ (grey bands marking the 1-$\sigma$ uncertainty).
Other points in Fig.~\ref{fig|DutyCycleHighz} come from other
studies with similar redshift ranges and luminosity thresholds.
At intermediate redshifts, the data of Bundy et al.\ (2008)
from the DEEP2 and AEGIS surveys show nearly the same
mass trend and normalization as Xue et al. (2010). More recently,
Mainieri et al. (2011) also found evidence for a steep
increase in AGN activity with stellar mass at $\langle z \rangle \approx 1.1$,
though with a luminosity threshold about two orders
of magnitude higher than the one by Xue et al. (2010).
At $z>2$ we add several points for high mass black holes
based on the studies of Caputi et al. (2006), Erb et al. (2006),
Kriek et al. (2007), and Alexander et al. (2005, 2008).
These data have been collected at bolometric
luminosities approximately $\log L > 43.8$, comparable to those of Xue et al. (2010).
Further details about these observations are discussed in
Appendix~\ref{Appendix|AGNdutyCycles}.
They extend the mass trend found by Xue et al. (2010), with
active fractions ranging from $\sim 10\%$ to $\sim 50\%$.
\subsection{Low-redshift AGN fractions: A redshift-dependent $P(\lambda)$?}\label{subsec|PLzdependence}
Red dotted and cyan dot-dashed curves in Fig.~\ref{fig|DutyCyclez0} show
the duty cycles $U(M_{\rm BH},z)$\ predicted by the G and G+P models at $z=0.2$,
where we have imposed the bolometric luminosity thresholds
$\log L > 41.5$ (left panel), $\log L > 43.5$ (middle), and
$\lambda > 0.01$ (right). For the right panel, we have additionally
multiplied model predictions by a mass-dependent factor
(ranging from $\sim$ 15\% at $M_{\rm BH} = 10^7 M_{\odot}$ to $\sim 3\%$ at $M_{\rm BH} = 3 \times 10^8 M_{\odot}$)
to account for Greene \& Ho's (2009) estimate of broad line AGN fractions.
(Specifically, the correction is obtained from the ratio of Broad Line to
Type II number densities
in their Figure 10, with an extrapolation of their results
above $M_{\rm BH} \sim 5 \times 10^8 M_{\odot}$.)
Both models drastically underpredict all measured active fractions for $M_{\rm BH}
> 10^{7.5} M_\odot$.
Predicted duty cycles are higher for the G+P model, and the gap
between G+P and G is larger for lower luminosity thresholds that allow
a larger contribution of sub-Eddington black holes, but the G+P model
still falls well short of the observed active fractions.
As introduced by Shankar et al. (2004) and discussed more
extensively by SWM (see their Figure 11), one can increase
predicted duty cycles at low redshift by adopting a redshift-dependent
Eddington ratio distribution, with characteristic $\lambda_c$
that drops towards low redshifts. Decreasing $\lambda_c \propto L/M_{\rm BH}$
maps a given luminosity to more massive, and hence rarer, black holes,
thus requiring a higher duty cycle to reproduce the observed AGN
space density. Furthermore, the results of H09,
KH09, and other studies directly suggest a
lower $\lambda_c$ at low redshifts, as already discussed in
Section~\ref{subsec|CompareWithObservedPL} (see Fig.~\ref{fig|CompareKauff1}).
Motivated by these results, we introduce models in which the location
of the peak in the G and G+P $P(\lambda)$\ distributions decreases
from $\lambda_c=1.0$ at $z=6$ to $\lambda_c \sim 0.02$ at $z\sim 0.1$,
following
\begin{equation}
\lambda_c(z)=\lambda(z=6)\left[\frac{1+z}{7} \right]^{\alpha}\, ,
\label{eq|PLz}
\end{equation}
with $\alpha=2.2$, although we note that even lower values of $\alpha$
yield similar results.
Eq.~\ref{eq|PLz} implies $\lambda_c$ = 0.71, 0.48, 0.29, 0.16, 0.06,
0.03, 0.02 at $z=5$, 4, 3, 2, 1, 0.5, and 0.1, respectively.
In agreement with SWM,
who considered $\delta$-function $P(\lambda)$, we find that this
evolution has a minor impact on the evolved black hole mass function (at
least at high masses) or on
the global accretion histories
shown in Fig.~\ref{fig|RhoBHz}. However,
predicted duty cycles at $z\sim 0.2$ for the G$(z)$\ model ---
a Gaussian $P(\lambda)$\ with the $\lambda_c(z)$ given by
Eq.~\ref{eq|PLz} --- are much higher than those of the
G model, as shown by comparing
the dotted and long-dashed curves in
Fig.~\ref{fig|DutyCyclez0}. The model now approximately agrees with the
Goulding et al. (2010), Kauffmann et al. (2003), and Schulze \& Wisotzki
(2010) active fractions,
though at high masses it is above the latter data and below
the former two.
Solid and short-dashed curves in Fig.~\ref{fig|DutyCyclez0} show
models with Gaussian and G+P $\lambda$-distributions
that incorporate both redshift and mass dependence of $\lambda_c$,
as discussed in the next section.
The redshift dependence expressed in Eq.~\ref{eq|PLz} is by
no means unique.
In particular, we have checked that an
equally good match to the local data can be obtained by allowing
the characteristic $\lambda_c$\
to decrease only below redshift $z\sim 1$, while remaining close to unity
at higher redshifts.
This abrupt change in $\lambda_c$\ at late times\footnote{We note that Shankar, Bernardi \& Haiman (2009) found that a constant
Eddington ratio as a function of redshift was consistent with the
statistics of local, early-type galaxies coupled with a mild evolution in the black hole-velocity dispersion scaling relation and the integrated AGN energy density.
However, their conclusions
were driven by assuming a non-evolving structural evolution of their hosts.
Allowing for some evolution in velocity dispersion (as suggested by several observations
and models, as in, e.g., Shankar et al. 2011, and references therein)
would be consistent with a decrease in time of the characteristic Eddington ratio
while preserving the mild evolution in the black hole-velocity dispersion relation.} leads to duty cycles that increase
at fixed black hole
mass from $z\sim 1$ to $z\sim 0$, instead
of decreasing as predicted by continuous redshift evolution.
This increasing behaviour may be in conflict with observations;
some direct X-ray
and optical analyses show that the fraction
of active galaxies above a fixed stellar mass decreases with decreasing
redshift in the range $0<z<1$ (e.g., Shi et al. 2008).
\subsection{AGN fractions vs. mass: A mass-dependent $P(\lambda)$?}
\label{subsec|PLmassdependence}
Fig.~\ref{fig|DutyCycleHighz} compares the model predictions to
the higher redshift active fraction data discussed in
Section~\ref{subsec|ObservationalEstimates}. The G and G$(z)$\
models both predict a duty cycle that declines with increasing $M_{\rm BH}$,
opposite to the trend found by
Xue et al.\ (2010) and Bundy et al. (2008).
At first glance, the G+P model appears to fare
better, at least in the $z=1.3$ and $z=2.3$ panels.
However, we view this better agreement as artificial ---
it is a consequence of the high initial black hole space
density required to keep high-$z$ duty cycles below
unity in this model (see Fig.~\ref{fig|BHMFandDuty}).
The high black hole space density inherited from these initial
conditions leads to a lower duty cycle for low mass black holes.
We regard these high initial space densities as observationally
untenable for the reasons discussed in Section~\ref{subsec|AccretionHistories},
so we discount this apparent success of the G+P model,
which vanishes by $z=0.5$ in any case.
\begin{figure*}
\includegraphics[width=15truecm]{Figure9bis.eps}
\caption{$P(\lambda)$\ distributions for our G$(z,M_{\rm BH})$\ model ({\it dashed} curves)
and the G$(z,M_{\rm BH})$\ model with a steep power-law component added
at low redshift ({\it solid} curves), shown at $z=0.6$ ({\it left})
and $z=0.2$ ({\it right}). In each panel, the three curves
show $P(\lambda)$\ for black hole masses of $10^7 M_\odot$,
$10^8 M_\odot$, and $10^9 M_\odot$ from top to bottom, as labelled.}
\label{fig|SteepPowerLaw}
\end{figure*}
The only other way we have found to make the predicted duty cycle
increase with black hole mass at these redshifts is to introduce
an explicit mass dependence of $P(\lambda)$, with $\lambda_c$ declining
with increasing mass. In this case massive black holes
are matched to more common, lower luminosity AGN, implying
a higher duty cycle. There is some direct empirical support for
such a trend (e.g. Heckman et al. 2004; Ballo et al. 2007;
Babi\'{c} et al. 2007; Netzer \& Trakhtenbrot 2007;
Rovilos \& Georgantopoulos 2007; Fine et al. 2008; H09),
and it could arise theoretically in models that envisage
a faster shut-off of activity in more massive systems due to AGN feedback
(e.g., Matteucci 1994; Granato et al. 2006; Lapi et al. 2006).
Motivated roughly by these empirical studies, we have adopted
a model with the same redshift dependence as before and a
mass dependence
\begin{equation}
\log \lambda_c(M_{\rm BH},z)=\log \lambda_c (z) +\left [\beta_M (\log M_{\rm BH}-C)
\right] \, ,
\label{eq|PLmassDependent}
\end{equation}
with $\beta_M=-0.3$ and $C=7$.
Eq.~\ref{eq|PLmassDependent} implies a $\lambda_c$ that drops by
a factor of two between $10^7 M_\odot$ and $10^8 M_\odot$ and
by a further factor of two between $10^8 M_\odot$ and $10^9 M_\odot$.
Solid lines in Figures~\ref{fig|DutyCyclez0} and~\ref{fig|DutyCycleHighz}
show predictions of the G$(z,M_{\rm BH})$\ model, which incorporates mass-dependence
of the radiative efficiency
$\epsilon$ as well as of $\lambda_c$, for reasons that we will
discuss shortly. This model also incorporates a broadening of the
Gaussian in the G$(z,M_{\rm BH})$\ at lower redshifts following the trend
\begin{equation}
\log \Sigma_\lambda(z)=\log \Sigma_\lambda(z=6)-0.4\log
\left(\frac{1+z}{7}\right) \, ,
\label{eq|Sigma}
\end{equation}
with $\log \Sigma_\lambda(z=6)=0.3$, yielding $\Sigma_\lambda(z) \sim 0.4,
0.5, 0.6$ at $z=3$, 2, 0.2.
As discussed below, we include this additional modification to the model to
provide a better match
to the local Eddington ratio distributions.
At all redshifts, this model predicts a duty
cycle that is roughly flat or rising
with black hole mass, though the rising trend is always
weaker than that found by Xue et al.\ (2010) and Bundy et al.\ (2008).
We could adopt a still stronger mass dependence in this model,
which would improve agreement with the observed trends,
but this would no longer be supported by empirical estimates
(e.g., Hickox et al.\ 2009), and it would exacerbate problems in
explaining the Eddington ratio distributions and local black hole mass function
(see below).
The G$(z,M_{\rm BH})$\ model is also in rough agreement with all
the local data sets on active fractions (solid lines in
Fig.~\ref{fig|DutyCyclez0}), though it overpredicts the Schulze \&
Wisotzki (2010) active fractions for $M_{\rm BH} \sim 10^9\, M_{\odot}$
by a factor of several.
Short-dashed lines in
Fig.~\ref{fig|DutyCycleHighz} show the predictions of
a model labeled G+P$(z,M_{\rm BH})$\ that has a G+P form with redshift-
and mass-dependent $\lambda_c$ and mass-dependent $\epsilon$
like that of the G$(z,M_{\rm BH})$\ model. Instead of allowing the Gaussian
to broaden at low redshift via Eq.~\ref{eq|Sigma}, we keep its
width fixed at $\Sigma_\lambda = 0.3$ but allow the power-law
tail to grow over time.
Specifically, the $P(\lambda)$ distribution is cut off at
\begin{equation}
\label{eq|LambdaMin}
\lambda_{\rm min}(z)= 0.1\left({1+z \over 7}\right)^3~,
\end{equation}
implying $\lambda_{\rm min} = $ 0.1, 0.063, 0.036, 0.019,
0.0079, 0.0023, 0.0098, 0.00039 at
$z = 6$, 5, 4, 3, 2, 1, 0.5, 0.1.
This simple modification (similar to that adopted by Cao 2010)
avoids the need for a large initial black hole seed
population, thus removing the artificial aspect of
our redshift-independent G+P model, but it produces
a full G+P distribution at intermediate and low redshifts.
The predictions of this model are similar to those of G$(z,M_{\rm BH})$,
with a somewhat worse match to the $z=1.3$ data in
Fig.~\ref{fig|DutyCycleHighz}b and a somewhat better
match to the Schulze \& Wisotzki data in
Fig.~\ref{fig|DutyCyclez0}. Both models fail to
reproduce the rising $U(M_{\rm BH})$ found at
$z=0.5$ by Xue et al. (2010).
\subsection{AGN fractions vs. mass: a steep power-law component at low $z$?}
\label{subsec|PLmassdependence2}
At early stages of our investigation, we anticipated that the observed
trends of rising active fraction with rising galaxy mass might be
explained mainly by the combination of a broad $P(\lambda)$ with
sample luminosity thresholds: high mass black holes remain observable
at lower $\lambda$ values, so they could have higher duty cycles
above fixed luminosity even if the trend of total duty cycle
(over all $\lambda$ values) was flat or decreasing.
This is essentially the explanation advanced by
Aird et al.\ (2011), who fit AGN data from the
PRIMUS survey with a model in which all black holes have
an Eddington ratio distribution $P(\lambda) = P_0(z)\times (\lambda/\lambda_0)^{-0.7}$
independent of mass.
As already discussed in the context of our
G+P model, we find that a long power-law tail is unrealistic
at high redshift because it implies duty cycles higher than unity
unless the population of seed black holes is implausibly large.
However, a power-law component that kicks in at lower redshifts
is still feasible. The blue dotted lines in
Figures~\ref{fig|DutyCyclez0} and \ref{fig|DutyCycleHighz}
show the predicted duty cycles for a model equivalent to
G$(z,M_{\rm BH})$\ with the addition
of a power-law component at $z < 0.7$ with slope $\alpha=-0.9$
and normalized to have the same value as the Gaussian at $\log \lambda=\log \lambda_c-0.2\log \Sigma_{\lambda}$.
Fig.~\ref{fig|SteepPowerLaw} plots $P(\lambda)$\ for this model at
three different black hole masses and two redshifts, in comparison
to those of the G$(z,M_{\rm BH})$\ model. For this plot we have multiplied $P(\lambda)$\
by $U(M_{\rm BH},z)$\ so that the curves are normalized to the duty cycle rather
than integrating to unit probability. The Gaussian and power-law
components join to produce a $P(\lambda)$\ that is close to a single
power law from $\lambda = 10^{-3}$ to $\lambda=1$.
This model still exhibits ``downsizing'' in the sense that higher
mass black holes have lower duty cycle at any given $\lambda$.
However, the duty cycle for AGN active above a luminosity
threshold $\log L = 43$ {\it increases} with black hole mass,
as shown in the left panel of Fig.~\ref{fig|DutyCycleHighz},
because massive black holes can shine above the threshold
at low $\lambda$. This is the only model we have constructed
that reproduces the Xue et al.\ (2010) trend at $z=0.5$.
The Aird et al.\ (2011) prescription, with the $P(\lambda)$\ normalization
independent of mass, would produce a still stronger rising trend,
but we find that such a model cannot simultaneously match our
evolved black hole mass function and our input luminosity function.
\section{Other observational constraints}
\label{sec|OtherConstraints}
\subsection{The local mass function: A mass-dependent radiative efficiency?}
\label{subsec|epsilonmass}
\begin{figure*}
\includegraphics[width=15truecm]{Figure9.eps}
\caption{Local black hole mass function
predicted by different models, as labelled.
Data are as in Fig.~\ref{fig|BHMFandDuty}.
Of the three G$(z,M_{\rm BH})$\ models, with mass-dependent $\lambda_c$,
only the one with mass-dependent $\epsilon$ (solid line) matches
the high end of the mass function. Elsewhere in the paper
this preferred model is simply labeled G$(z,M_{\rm BH})$.}
\label{fig|BHMFcomparison}
\end{figure*}
As shown previously in Fig.~\ref{fig|BHMFandDuty}, our simple, non-evolving
$\delta$, G, and G+P models yield reasonable agreement with
SWM's estimate of the local black hole mass function for
$M_{\rm BH} \leq 10^9 M_\odot$, but the model predictions skirt
the upper boundary of the observational estimates at higher
masses. Fig.~\ref{fig|BHMFcomparison} plots mass functions
for the Gaussian $P(\lambda)$\ models with various assumptions about
redshift evolution and mass-dependence.
Redshift evolution alone (red long-dashed line) produces modest changes above
$\sim 2 \times 10^8 M_\odot$, but the decreasing $\lambda_c$ boosts growth
of high mass black holes at the expense of low mass black
holes, leading to a flattening of $\Phi(M_{\rm BH})$ near
$10^8 M_\odot$ and a lower space density of lower mass black holes.
This change partly reproduces the turnover found by
Vika et al. (2009), though it still does not reproduce
their estimate below $M_{\rm BH} \sim 10^{7.5} M_\odot$.
Adding only the mass-dependent $\lambda_c$ of Eq.~\ref{eq|PLmassDependent}
leads to the ``G$(z,M_{\rm BH})$, constant $\epsilon$'' curve (blue dot-dashed line)
in Fig.~\ref{fig|BHMFcomparison}.
For this model, we have lowered the radiative efficiency from
$\epsilon = 0.06$ to $\epsilon = 0.05$ to improve the match to the amplitude
of $\Phi(M_{\rm BH})$ at $10^8-10^{8.5} M_\odot$, where it is
best measured. However, the model then strongly overpredicts
the mass function at $M_{\rm BH} > 10^9 M_\odot$, a direct
consequence of the higher duty cycles of high mass black holes
that the model was designed to produce. As discussed in
Section~\ref{subsec|Mergers} below, the expected impact of black
hole mergers would make this overprediction even more severe.
Volonteri, Sikora \& Lasota (2007), Cao \& Li (2008), and Fanidakis et
al. (2011), among others, have suggested that
radiative efficiency may
increase with black hole mass, perhaps mirroring a merger-induced
increase in the average spin for the more massive black
holes. Cao \& Li (2008) claimed evidence for an increasing
$\epsilon$\ from the match between the predicted and local black
hole mass functions at the high-mass end. More recently
Davis \& Laor (2011) directly determined the radiative efficiency
from the ratio between bolometric luminosity and accretion rate, the latter
determined from thin accretion disk model fits to the optical luminosity
density.
Their analysis seems to support an increase of the radiative efficiency
with black hole mass, from 0.03 at low masses to 0.4 at high masses.
An increasing radiative efficiency at high masses
is also consistent with the notion that the bulk of luminous radio
AGNs, believed
to be rapidly spinning black holes, are indeed massive black holes (e.g.,
McLure \& Jarvis 2004; Metcalf \& Magliocchetti 2006;
Shankar et al. 2008a,b; Shankar et al. 2010e).
Inspired by these theoretical and empirical arguments,
we have constructed the G$(z,M_{\rm BH})$\ model (see Tables~\ref{table|models} and \ref{table|models2}), which, following Cao \& Li (2008), adopts a
mass-dependent $\epsilon$\ given by:
\begin{equation}
\epsilon=\left\{
\begin{array}{ll}
0.05 & \hbox{if $M_{\rm BH} < 10^8 $$M_{\odot}$} \\
0.05(M_{\rm BH}/10^8\, M_{\odot})^{0.3} & \hbox{if $M_{\rm BH} \ge 10^8
$$M_{\odot}$\,\, .}
\end{array}
\right. \label{eq|radefficiencyMass}
\end{equation}
This change reduces the implied accretion rates of massive black
holes, leading to much better agreement with the local
$\Phi(M_{\rm BH})$ as shown by the solid curve in
Fig.~\ref{fig|BHMFcomparison}. This model also has a stronger
positive trend of duty cycle with $M_{\rm BH}$, improving agreement
with the data in Fig.~\ref{fig|DutyCycleHighz}.
The two improvements are connected: raising $\epsilon$ decreases
the space density of massive black holes, so a higher duty cycle
is required to match the space density of active black holes.
The mass function (not shown) of the model introduced in
Section~\ref{subsec|PLmassdependence2}, with a steep power-law
at $z<0.7$, is nearly identical to that of G$(z,M_{\rm BH})$, except at
$\log M \leq 7.2$, where it is lower by $\sim 0.3$ dex.
Wang et al. (2009) have instead recently claimed empirical
evidence for an increase of $\epsilon$\ with redshift.
Their results are based on an inversion of Soltan's (1982) argument,
expressing the radiative efficiency at any redshift
as the ratio of the accreted mass density up to that redshift
$\rho_{\rm BH}(z)$ to the
corresponding total emissivity obtained by direct
integration of the AGN luminosity function (see their Eq.~6).
The mass density $\rho_{\rm BH}(z)$ was computed by first measuring the
mass density locked up in all active
black holes at $z$ (with masses from virial relations), then correcting
by the duty
cycle extracted from the number counts of active galaxies in VIMOS-VLT
Deep Surveys.
Clustering analysis also provides hints of redshift-dependent
radiative efficiency.
Shankar et al. (2010b), adopting basic
accretion models and cumulative number matching
arguments, found that black hole accretion plus merger models consistent
with both the quasar luminosity function and the
strong observed clustering at $z \approx 4$ (Shen et al. 2007)
must be characterized by high duty
cycles and large radiative efficiencies $\epsilon \gtrsim 0.2$, if they
are accreting at a significant fraction of the Eddington limit.
However, an efficiency $\epsilon \geq 0.2$ at all redshifts
would underpredict the local black hole mass function (SWM).
In flux-limited quasar surveys, higher redshift quasars are found
above higher luminosity thresholds, so
both of these empirical arguments could potentially be answered by a
mass-dependence of the sort implied by Eq.~\ref{eq|PLmassDependent}.
We nonetheless consider a directly redshift-dependent model,
approximating these empirical findings with the relation
\begin{equation}
\epsilon(z)=0.0022\left[1+{\rm erfc} \left(-\frac{z}{1.5}\right)^6\right]\, .
\label{eq|radEfficiencyReds}
\end{equation}
Eq.~(\ref{eq|radEfficiencyReds}) implies
$\epsilon \approx 0.14$, 0.14, 0.12, 0.05, 0.02, 0.01
at $z=6$, 4, 2, 1, 0.5, 0.2, respectively.
The local black hole mass function predicted by this G$(z,M_{\rm BH})+\epsilon(z)$
model,
shown by a green dashed curve in Fig.~\ref{fig|BHMFcomparison},
tends to overestimate the observed mass function at all scales.
The match could be improved
by increasing the overall normalization of $\epsilon(z)$, still
allowed (and actually preferred) by the measurements of Wang et al. (2009;
see their Figure 2\emph{a}), but such an increase
produces unphysical models with duty cycles significantly
higher than unity, and it still leaves too many high mass black holes.
Moreover, the cosmological accretion rate predicted by the
G$(z,M_{\rm BH})+\epsilon(z)$ model is morphologically different from the
cosmological star formation rate of galaxies, at variance
with the agreement shown in Fig.~\ref{fig|RhoBHz}.
In itself this is not a fatal objection, but it then requires
non-trivial fine tuning to reproduce a
tight local relation between black hole mass and stellar mass.
\subsection{Eddington Ratio Distributions, Revisited}
\label{subsec|EddingtonRatioDistributions}
\begin{figure*}
\includegraphics[width=15truecm]{Figure10.eps}
\caption{Contributions of different Eddington ratio distributions
to the overall bolometric luminosity functions at $z=3.2$ (top) and $z=0.3$
(bottom),
as predicted by the G$(z)$, G$(z,M_{\rm BH})$, and G+P$(z,M_{\rm BH})$\ models (left, middle, right
panels, respectively).}
\label{fig|FLlambdaContributions}
\end{figure*}
\begin{figure*}
\includegraphics[width=15truecm]{Figure11.eps}
\caption{Similar format to Figures~\ref{fig|CompareKollmeier1} and
\ref{fig|CompareKauff1}. Comparison among the G$(z,M_{\rm BH})$, G$(z)$\ and
G+P$(z,M_{\rm BH})$\ predicted Eddington ratio distributions (\emph{solid}, \emph{red
long-dashed}, and \emph{cyan dashed} lines, respectively) compared to the
K06 data at $z=2$ and $z=1$, respectively (\emph{top}
and \emph{middle} panels), and to H09 at $z=0.5$
(\emph{lower left}) and KH09 in the local
Universe (\emph{lower right}). In the lower panels, we also include
curves for the steep power-law model introduced in
Section~\ref{subsec|PLmassdependence2}.
}
\label{fig|CompareKollKauff}
\end{figure*}
In any model with a broad $P(\lambda)$,
the observable distribution of Eddington ratios depends on the
luminosity of the AGN being considered. Fig.~\ref{fig|FLlambdaContributions}
shows the contribution to the bolometric luminosity function
from different ranges of $\lambda$ at $z=3.2$ (top panels) and $z=0.3$
(bottom panels) in the
G$(z)$, G$(z,M_{\rm BH})$, and G+P$(z,M_{\rm BH})$$\,$ models (left, middle, and right panels, respectively).
There is a natural trend for the high end of the
luminosity function ($\log L > 46$) to be mainly contributed by high-$\lambda$
black holes, with a broader range of $\lambda$
contributing at lower luminosities.
Interestingly, we find that the $\lambda < 0.01$ range does not
dominate at any luminosity at any redshift.
Only in the G$(z,M_{\rm BH})$\ model and at low redshifts are there roughly equal
contributions to the luminosity function from each
logarithmic bin of $\lambda$ in the range $0.01-1.0$.
This is important to compare with models that rely on
a significant contribution from ``ADAF-type'' modes
to the global accretion history of black holes
(see, e.g., Merloni \& Heinz 2008 and Draper \& Ballantyne 2010).
Fig.~\ref{fig|CompareKollKauff} compares the Eddington ratio
distributions of the G$(z,M_{\rm BH})$, G$(z)$, and G+P$(z,M_{\rm BH})$\ models (black solid, red long-dashed,
and cyan dashed lines, respectively) to the observational estimates of K06,
H09, and KH09,
discussed earlier in Section~\ref{subsec|CompareWithObservedPL}.
Beginning at low redshift (bottom panels), we see that the
declining $\lambda_c(z)$ in the redshift-dependent models
resolves the discrepancy seen in Fig.~\ref{fig|CompareKauff1},
producing much better agreement with the H09
and KH09 data. The additional broadening in the Gaussian component
(Eq.~\ref{eq|Sigma})
produces excellent agreement with the H09 distribution
and excellent agreement with KH09 for $\lambda > 10^{-2.5}$.
The G+P$(z,M_{\rm BH})$\ model achieves similar agreement, matching KH09
slightly better at low luminosities but still falling short
at $\lambda < 10^{-3}$.
The blue dotted curves in the lower panels show the G$(z,M_{\rm BH})$\ model with
the steep power-law at low $z$, introduced in
Section~\ref{subsec|PLmassdependence2}. The predicted
Eddington ratio agrees poorly with both the H09 and KH09 data.
At $z>1.2$ (upper panels) model predictions are in approximate agreement
the K06 data, with G$(z)$\ showing the best agreement.
However, the $z=1$ model outputs tend to disagree with the K06, $z<1.2$
histograms (middle panels).
We note that Netzer \& Trakhtenbrot (2007) find a
peak at $\lambda \approx 0.1$ for $M=10^{8}-10^{8.5} M_\odot$
black holes at $z=0.7$. Kelly et al. (2010) have also recently claimed an Eddington ratio distribution
from SDSS of Broad Line Quasars that peaks at $L/L_{\rm Edd} \sim 0.05$
with a dispersion of $\sim 0.4$ dex.
Both these results would be closer to the model predictions.
While the duty cycle data considered in Section~\ref{subsec|PLmassdependence}
seem to favor a $\lambda_c$ that decreases at higher $M_{\rm BH}$,
the K06 data disfavor this solution.
We note that the luminosity threshold of the Xue et al. (2010)
study that motivates this model is much lower than that of the
K06 data, $L_{\rm min} \sim 10^{43}{\rm erg \, s^{-1}}$ rather
than $L_{\rm min} \sim 10^{45}-10^{46}{\rm erg \, s^{-1}}$. A reconciliation
of these results could therefore lie in a model that behaves
differently in these two luminosity regimes.
\subsection{The Impact Of Mergers}
\label{subsec|Mergers}
\begin{figure*}
\includegraphics[width=10truecm]{Figure12.eps}
\caption{Predicted black hole mass function for the G$(z,M_{\rm BH})$\ model
without mergers (\emph{dashed} lines), with only major mergers
(\emph{solid} lines),
and with minor and major mergers (\emph{dotted} lines). The
set of lines, from bottom to top, are the model predictions at $z=4, 2,
0.1$, respectively. The data are as in Fig.~\ref{fig|BHMFandDuty}. The
cumulative effect of mergers is minor at $z>2$ but becomes
significant at low redshifts and high masses.}
\label{fig|Mergers}
\end{figure*}
It is usually assumed that mergers of galaxies are followed by mergers
of their central black holes, making black hole mergers a potentially
important source of evolution in the black hole mass function
(e.g., Hughes \& Blandford 2003; Wyithe \& Loeb 2003;
Islam et al. 2004; Scannapieco \& Oh 2004; Yoo \& Miralda-Escud\'{e} 2004;
Volonteri et al. 2005; Lapi et al. 2006; Yoo et al. 2007; Marulli et al. 2008;
Bonoli et al. 2009; Shen 2009; SWM; Bonoli et al. 2010; Shankar 2010; Shankar et al. 2010a,d; Shankar et al. 2011;
Kocsis \& Sesana 2011; Kulkarni \& Loeb 2011).
We will save a full discussion of mergers for future work, but here
we briefly assess their potential impact on the mass function.
Improving on our simplified calculation in SWM, we here follow the schemes
proposed by Shen (2009) and Shankar et al. (2010b), computing the
rate of black hole mergers from the halo merger rate
predicted from fits to N-body dark matter simulations (Fakhouri \& Ma 2008),
corrected by a dynamical friction timescale.
In the presence of mergers
the continuity equation reads as
\begin{equation}
\frac{\partial n_{\rm BH}}{\partial
t}(M_{\rm BH},t)=-\frac{\partial (\langle \dot{M}_{\rm BH}\rangle
n_{\rm BH}(M_{\rm BH},t))}{\partial M_{\rm BH}}+ S_{\rm in} - S_{\rm out}\,,
\label{eqApp|contEqMergers}
\end{equation}
where $S_{\rm in}$ and $S_{\rm out}$
are, respectively, the merger rate of smaller mass black holes ending
up with mass $M_{\rm BH}$\ and the merger rate of black holes with initial mass $M_{\rm BH}$\
merging into more massive systems.
The merger rate of haloes as a function of redshift, mass ratio of the progenitors, and
mass of the remnants are taken from Fakhouri \& Ma (2008).
We then convert the merger rate of haloes to a merger rate of black holes
via the median $M_{\rm BH}$-$M_{\rm halo}$\ relation,
defined, at all times, by cumulative number matching between
the black hole and halo mass functions,
which allows us to associate the proper halo
merger rate to a given bin of black hole mass.
Full details are given in
Appendices~\ref{Appendix|MbhMhaloRelation} and~\ref{Appendix|Mergers}.
We make the limiting case assumption that halo mergers are always
followed by black hole mergers after a dynamical friction time.
The accuracy of this assumption remains a matter
of debate (e.g., Cavaliere \& Vittorini 2000; Shen 2009; Shankar 2010,
and references therein). The impact of the dynamical friction time delay
itself is irrelevant at $z \lesssim 2$ (see Shen 2009).
Fig.~\ref{fig|Mergers} shows the mass function of the G$(z,M_{\rm BH})$\ model
at $z=4$, 2, and 0.1 without mergers (dashed lines), including major
mergers
above a black hole mass-ratio threshold $\xi=0.5$ (solid lines), and including
all mergers above a black hole mass-ratio threshold $\xi=0.1$ (dotted lines).
Mergers have limited effect at $z > 2$, but by $z=0.1$ they
have dramatically boosted the space density of black holes with
$M_{\rm BH} > 10^9 M_\odot$, to a level clearly inconsistent with the
SWM and Vika et al. (2009) estimates.
The discrepancy would be more severe if we did not include
mass-dependent $\epsilon$ in the G$(z,M_{\rm BH})$\ model
(see Fig.~\ref{fig|BHMFcomparison}).
We caution, however, that
in this regime the local mass function estimates rely largely on
extrapolation of the $M_{\rm BH}-\sigma_*$ or $M_{\rm BH}-M_*$ correlations
into a range with limited observational constraints. This is also
a mass range where typical galaxy hosts are gas poor, and it is
not clear that ``dry'' galaxy mergers necessarily lead to black
hole mergers.
In general, the inclusion of mergers makes it even harder
to fit the observed steep decline of black hole abundance at high masses.
\subsection{A $\lambda$-dependent bolometric correction}
\label{subsec|kbolLambda}
The inclusion of mergers puts otherwise acceptable models
at risk of overpredicting the high mass end of the local black
hole mass function. We now discuss one possible resolution of
this tension, a $\lambda$-dependent bolometric correction.
\begin{figure}
\includegraphics[width=8.5truecm]{Figure14.eps}
\caption{Binned data on the $2-10$ keV X-ray bolometric correction as
a function of Eddington
ratio from Vasudevan \& Fabian (2007). The \emph{solid} line is the analytical
approximation used
in this paper, while the \emph{long-dashed} and \emph{dot-dashed} lines
are the results from Lusso et al. (2010).}
\label{fig|KbolLambda}
\end{figure}
Following previous work
(e.g., Elvis et al. 1994; Marconi et al. 2004; Shankar et al. 2004;
Hopkins et al. 2007), we have so far assumed that the bolometric
correction depends only on bolometric luminosity (all the X-ray data
in SWM were converted to bolometric luminosities using the
$L$-dependent bolometric correction by Marconi et al. 2004). However, some
studies (e.g., Dai et al. 2004; Saez et al. 2008; and references
therein) show signs for variations in the spectral energy
distributions of AGNs with either redshift or accretion properties.
In particular, Vasudevan \& Fabian (2007, 2009) suggest that variations in
the disc emission in the ultraviolet may be important to build the
optical-to-X-ray spectral energy distributions of AGNs. From a
sample of 54 AGNs from the Far Ultraviolet Spectroscopic Explorer
(FUSE) and X-ray data from the literature, they claim evidence
for a large spread in the bolometric corrections, with no
simple dependence on luminosity being evident. Their results suggest
instead a more well-defined relationship between the bolometric correction
and Eddington ratio, with a transitional region at an Eddington
ratio of $\sim 0.1$, below which the X-ray bolometric correction is
typically 15-25, and above which it is typically 40-70. As shown in
Fig.~\ref{fig|KbolLambda}, we approximate their results by setting
\begin{equation}
K_{\rm 2-10 keV}(\lambda)= \left\{
\begin{array}{lll}
18 & \hbox{if $\lambda \le 0.105$} \\
54.85+26.78\, \log L_X-\\
-11.11 \, (\log
L_X)^2 & \hbox{if $0.105\le \lambda \le 1$\,\, .}
\end{array}
\right. \label{eq|BClambda}
\end{equation}
Our analytic approximation is shown as a solid line in
Fig.~\ref{fig|KbolLambda},
against the binned data by Vasudevan and Fabian (2007; the fit and the data
extend to $\lambda>1$), while the
dot-dashed and long-dashed lines are more recent fits from Lusso et al. (2010)
derived
from a larger sample. As the latter authors point out, while a trend
of bolometric correction with $\lambda$\ might exist, determining its exact
slope is still challenging given the large dispersion in the data. In this
section we will use Eq.~\ref{eq|BClambda} as a reference, noting that
the Lusso et al. (2010) fits
provide consistent results in the range of interest here.
\begin{figure}
\includegraphics[width=8.5truecm]{Figure15.eps}
\caption{Influence of a $\lambda$-dependent bolometric correction
on the predicted black hole mass function. The {\emph dot-dashed} line
shows the $z=0$ mass function of the
G$(z,M_{\rm BH})$\ model with our standard bolometric correction.
The {\emph solid} lines show the evolving mass function at $z=2$, 1, and 0
for the same model assuming the $\lambda$-dependent bolometric
correction of Eq.~\ref{eq|BClambda}. The dotted line is the $\lambda$-dependent G$(z,M_{\rm BH})$\ model inclusive of mergers (with $\xi >0.3$). Data points and grey band are the same
as those in Figs.~\ref{fig|BHMFandDuty} and~\ref{fig|BHMFcomparison}.}
\label{fig|NMKbolLambda}
\end{figure}
In our numerical formalism described in
Appendix~\ref{app|solvingContEq}, it is
straightforward to insert a $\lambda$-dependent bolometric correction. Given that
the relation $L\propto K_{\rm 2-10 keV}(\lambda) \times L_X \propto \lambda
M_{\rm BH}$, it implies that $L_X\propto [\lambda/K_{\rm 2-10 keV}(\lambda)] M_{\rm BH}
$. Therefore, having a $\lambda$-dependent bolometric correction is equivalent
to running the code replacing bolometric luminosities with
$L_X$. We thus solve Eq.~\ref{eq|PhiLLambda} for computing the duty cycle
by replacing the bolometric luminosity function
$\Phi(L,z)$ on the left-hand side with the X-ray luminosity function $\Phi_X(L_X,z)$,
and using effective Eddington ratios $\lambda'=\lambda/K_{\rm 2-10 keV}(\lambda)$.
Most predictions of our models are not sensitive to this change
of bolometric correction. In particular, we have
checked that after taking into account the $\lambda$-dependent
conversions between sample flux limits and bolometric
luminosities, both duty cycles and Eddington
ratio distributions showed similar behaviours
to the ones predicted by the G$(z,M_{\rm BH})$\ model with a luminosity-dependent
bolometric correction.
Nevertheless, the $\lambda$-dependent
correction does have a significant effect on the low-$z$ black
hole mass function.
The solid lines (at $z=0,1,2$, from top to bottom)
in Fig.~\ref{fig|NMKbolLambda} show the G$(z,M_{\rm BH})$\ model with a $\lambda$-dependent bolometric correction as in Eq.~\ref{eq|BClambda}.
High-mass black holes have preferentially lower $\lambda$ in this model,
and the lower bolometric luminosity at a given X-ray luminosity
reduces the inferred growth rate of these massive black holes.
Since G$(z,M_{\rm BH})$ previously agreed well with the $z=0$ mass function,
it now underpredicts the high mass end.
A $\lambda$-dependent bolometric correction would reduce the need
for mass-dependent $\epsilon$ in our G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$\ models
(Section~\ref{subsec|epsilonmass}, Fig.~\ref{fig|BHMFcomparison}), which
was inferred partly from the high mass end of the local mass
function, though it was also supported by
direct empirical evidence for mass-dependent $\epsilon$.
Alternatively, a $\lambda$-dependent
bolometric correction could compensate the impact of mergers on our
standard G$(z,M_{\rm BH})$\ model (see Fig.~\ref{fig|Mergers}) to yield improved
agreement with local mass function estimates. The dotted line
in Fig.~\ref{fig|NMKbolLambda} is the predicted $z=0$ black hole mass
function for the G$(z,M_{\rm BH})$\ model including a mass-dependent $\epsilon$,
$\lambda$-dependent bolometric correction, and black hole mergers
with $\xi > 0.3$. Agreement with observational estimates is
significantly improved relative to the standard bolometric
correction case (uppermost solid line in Fig.~\ref{fig|BHMFcomparison}).
Uncertainties in bolometric corrections --- their normalization
and their dependence on luminosity, $\lambda$, or other factors --- remain an important source of uncertainty when
testing evolutionary models of the black hole population
against the local census of black holes.
\subsection{Specific Black Hole Accretion Rate}
\label{subsec|SBHAR}
\begin{figure*}
\includegraphics[width=15truecm]{Figure16.eps}
\caption{Mean specific black hole accretion rate
(Eq.~\ref{eq|SBHAR})
predicted by
the G$(z)$\ (\emph{left}) and G$(z,M_{\rm BH})$\ (\emph{right}) models as
a function of black hole mass (\emph{top}) and of redshift
(\emph{bottom}). In the top panels,
\emph{cyan} lines are the mean specific star formation rates as calibrated
by Karim et al. (2011; their Table 4), with stellar masses simply scaled
to black hole masses assuming a proportionality factor of $10^{-3}$, as
measured in the local Universe. In the bottom panels,
the \emph{grey area} is the specific star
formation rate with error bars calibrated by Gonz{\'a}lez et al. (2010)
for galaxies with stellar mass $\sim 5\times 10^9 - 10^{10} M_{\odot}$,
which can be compared to the predictions for
$10^7M_{\odot}$ black holes (\emph{solid} curves).}
\label{fig|SBHAR}
\end{figure*}
Several authors have noted
that the average black hole accretion rate
has a redshift dependence morphologically similar
to the cosmological SFR (e.g., Marconi et al. 2004; Merloni et al. 2004;
Silverman et al. 2008a; Zheng et al. 2009; SWM).
Fig.~\ref{fig|SBHAR} adds a new piece of information to
the co-evolution of black holes and galaxies, plotting the mean specific
accretion rate of black holes of a given mass at a given redshift
defined as
\begin{equation}
\label{eq|SBHAR}
\frac{\langle \dot{M}(M_{\rm BH},z) \rangle}{M_{\rm BH}} = \frac{\langle \lambda(M_{\rm BH},z)
\rangle U(M_{\rm BH},z)}{t_s} \, .
\end{equation}
While the mean accretion rate
does not depend on the $P(\lambda)$\ distribution
but only on the assumed radiative efficiency,
the mean specific accretion rate depends significantly on the input $P(\lambda)$,
so it provides diagnostic power beyond that in the global rates.
In the G$(z)$\ model (left panels), the specific mean accretion rate declines with
increasing black hole mass at all epochs.
Cyan lines in the upper panels show the mean specific star
formation rate (SSFR) recently derived by Karim et al. (2011; see also Noeske
et al. 2009) at the
same redshifts (from their Table 4),
with stellar masses simply re-scaled by a factor of $10^{-3}$ to convert to
black hole masses.
In this simple comparison with the G$(z)$\ model,
the mean specific black hole accretion
rate decreases with mass and increases with redshift in a remarkably
similar way as the mean star formation rate, with a slope
$\sim M_{\rm BH}^{-0.4}$.
However, when we consider the G$(z,M_{\rm BH})$\ model (right panels), which better matches
data on black hole duty cycles and $\lambda$-distributions,
we predict higher specific accretion rates at high masses,
reflecting the higher duty cycles in this model.
For simplicity, here we use a constant radiative efficiency,
thus a constant $t_s$ in Eq.~\ref{eq|SBHAR}.
Silverman et al. (2009) also find that in relatively massive
high-$z$ galaxies the ratio between average black hole accretion rate and SFR
is higher by up to an order of magnitude with respect to the classical
$10^{-3}$, in fair agreement with the prediction of the G$(z,M_{\rm BH})$\ model,
though dependent on the actual variations of radiative efficiency with mass and/or time. In fact, we checked that including a mass-dependent radiative efficiency as in Eq.~\ref{eq|radefficiencyMass} would line up the specific black hole accretion rate with
the SSFR at all masses.
The lower panels of Fig.~\ref{fig|SBHAR}
show the mean specific black hole accretion rate
as a function of redshift for different bins of black hole mass, as labelled.
The grey bands indicate the uncertainties around the measured SSFR as
catalogued and derived by Gonz{\'a}lez et al. (2010) for galaxies with
stellar mass $\sim 5\times 10^9 - 10^{10} M_{\odot}$.
The latter should be compared with only the specific accretion rate onto
black holes with current mass $\sim 10^7 M_{\odot}$ (solid lines), but for
completeness
we also show the accretion rate for more massive black holes
$M_{\rm BH} = 10^8 M_{\odot}$ and $M_{\rm BH} = 10^9 M_{\odot}$ (long-dashed and dotted
lines, respectively), as labelled.
Overall, consistently with what is found in the upper panels,
the models predict an accretion rate that tracks the SSFR,
though the G$(z)$\ model tends to produce a specific accretion rate that
steadily increases even at $z \gtrsim 2$, at variance with the data for
the star formation rate.
The G$(z,M_{\rm BH})$\ model predicts a redshift dependence of the accretion rate
morphologically similar
to that of galaxies, a fact that might play some role
in explaining the still puzzling plateau at $z \gtrsim 2$ of the galactic
SSFR$(z)$, which is poorly reproduced by semi-analytic models and might
require some extra source of early feedback (e.g., Weinmann et al. 2011).
\section{Discussion}\label{sec|discu}
Although we have considered a wide variety of models, with different
$P(\lambda)$\ shapes and different redshift and mass dependences of $\lambda_c$
and $\epsilon$, every one of these models shows significant (factor
of several) disagreement
with at least one of the observational tests we have examined.
This failure could indicate that our model assumptions are still
too restrictive to describe the real black hole population --- for
example, we generally assume that the shape of $P(\lambda)$\ is independent of redshift
and black hole mass except for overall shifts in $\lambda_c$, and
we have considered restricted functional forms for these $\lambda_c$ trends
and for $P(\lambda)$\ itself. Alternatively, the problem could lie in one
or more of the data sets themselves, since these are frequently
derived from noisy or uncertain estimators (e.g., for black hole
masses) and from input samples that are subject to selection
biases and incompleteness (e.g., for active galaxy fractions).
Here we highlight aspects of the data that appear especially
difficult to reproduce within our class of models, or where different
data sets appear to drive the models in contradictory directions.
Recall that all of our models reproduce SWM's estimate of the
bolometric luminosity function (summarized in Section~\ref{subsec|AGNLF})
by construction. SWM discuss remaining uncertainties in this
luminosity function and their impact on inferred model parameters.
The first tension within the data is at the low-mass end of the local
black hole mass function (see Fig.~\ref{fig|BHMFandDuty}), where SWM
(and a number of other studies) find a $\Phi(M_{\rm BH})$ rising
to low masses but Vika et al. (2009) and some other studies
(e.g., Graham et al. 2007) find a falling $\Phi(M_{\rm BH})$. This region of
the mass
function remains difficult to probe because of uncertainties
in bulge-disk decomposition and because
the black hole mass correlations for spiral bulges
are more uncertain than those for high mass, bulge-dominated galaxies.
In our models, it is difficult to produce a turnover like that
of Vika et al. (2009), though the G$(z)$\ model goes in this direction
(Fig.~\ref{fig|BHMFcomparison}) because it ascribes much of the low
luminosity AGN activity to low-$\lambda$ accretion by massive
black holes, hence reducing the growth of low mass black holes.
However, the predicted mass function in this regime depends on the
input luminosity function in a range that is largely extrapolated
from brighter magnitudes. Therefore if $\Phi(M_{\rm BH})$ really does
turn over at low masses, a plausible explanation is that the faint
end of the AGN luminosity function is flatter than the one adopted here.
A second tension, with more serious implications for the issues
at the core of this paper, is between the narrow Eddington ratio
distributions measured by K06 and the broader
distribution, peaking at lower $\lambda_c$, measured by H09.
The two data sets overlap in redshift, though H09
are at the low redshift end of the K06 range.
H09, and KH09 at still lower $z$, favor a
$P(\lambda)$\ that is broad in shape (like G+P) and evolving to low $\lambda_c$
at low redshift, but it is difficult to reconcile such a model with
the K06 histograms. The luminosity thresholds
for the H09 and K06 data sets are very different, roughly
$L_{\rm min} = 10^{43}{\rm erg \, s^{-1}}$ and $L_{\rm min}=10^{45}{\rm erg \, s^{-1}}$,
respectively, so there is no direct contradiction between
the measurements. Possibly a model that allows a different
{\it shape} of $P(\lambda)$\ for high and low mass black holes, or that
allows a power-law that is steeper or further offset from the
log-normal peak, could be made consistent with both sets of
observations. The two samples might differ
for more profound physical reasons, as recently emphasized by
Trump et al. (2011), who showed that broad lines
might disappear at $\lambda < 0.01$ because of a change in
accretion flow structure.
The observed $P(\lambda)$\ distributions also include errors
in black hole mass estimations, so the intrinsic distributions
should in principle be even narrower.
Correcting for this observational broadening would exacerbate
the tension with our G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$\ models, which already
predict broader $P(\lambda)$\ distributions than those found by K06.
Another important tension arises between the low duty cycles found
for low redshift quasars by Schulze \& Wisotzki (2010) and the much
higher active galaxy fractions found by Kauffmann et al. (2003),
with a gap that is roughly two orders of magnitude. The difference
between Type 1 and Type 2 AGN could possibly explain a factor of $10-20$
(Greene \& Ho 2009), though this factor is already
large compared to conventional estimates of obscured-to-unobscured
AGN ratios, and other empirical and physical effects might
need to be invoked to explain
such a strong discrepancy (e.g., Trump et al. 2011).
By adopting the Greene \& Ho (2009) ratios in our predictions, we
find models that are roughly consistent with
both Kauffmann et al.\ (2003) and
Schulze \& Wisotzki (2010), but even our best cases disagree
with one of these data sets by a factor of several in some
black hole mass range (see Fig.~\ref{fig|DutyCyclez0}).
We have converted Kauffmann et al.'s [OIII] luminosities
to bolometric luminosities assuming a constant bolometric correction,
and scatter or biases in this correction
(see, e.g., Capetti\ 2011) might account for some of the
discrepancy with Schulze \& Wisotzki (2010).
Best et al.\ (2005, see their Fig.~2) also find high AGN
fractions for SDSS galaxies --- $20-40\%$ for
$L_{\rm [OIII]} > 10^{5.5}L_\odot$ ($\log L \ga 42$) --- comparable
to those of Goulding et al.\ (2010).
A fourth tension arises from the trend of higher active fractions
for more massive galaxies found by Xue et al. (2010) and
Bundy et al. (2010). If $P(\lambda)$\ does not evolve, then matching
observed AGN luminosity evolution leads to downsizing, i.e.,
a {\it decrease} of duty cycle with increasing black hole mass
(Fig.~\ref{fig|BHMFandDuty}). A $\lambda_c(z)$ that declines
towards low redshift can soften this trend, but within
our considered range of models, it does not eliminate downsizing
entirely. We have been able to produce a trend of rising duty
cycle with rising black hole mass by making $\lambda_c$ decrease
towards higher $M_{\rm BH}$, but the resulting models then tend to be inconsistent
with the K06 Eddington ratio distributions, which
do not show such a trend (Fig.~\ref{fig|CompareKollKauff}).
Furthermore, the model trends of $U(M_{\rm BH})$ remain
flatter than those found by Xue et al. (2010) and Bundy et al. (2008),
and these models still yield falling $U(M_{\rm BH})$ at $z=0.5$.
One caveat is that we are translating the observed galaxy stellar
masses to corresponding black hole masses assuming a linear relation
with no scatter. If the scatter between black
hole mass and galaxy mass were large at these redshifts, then
the trend between active fraction and galaxy mass could
be partly induced by a higher probability for black holes of fixed mass to
be active if they reside in more massive galaxies.
In Section~\ref{subsec|PLmassdependence2} we considered a model
in which $P(\lambda)$\ rises steeply towards low $\lambda$
(with $P\propto \lambda^{-0.9}$), which helps produce a rising $U(M_{\rm BH})$
trend for a luminosity thresholded sample because more massive black
holes can radiate at low $\lambda$ while remaining above
threshold (Aird et al. 2011; Mainieri et al. 2011).
We restricted this steep $P(\lambda)$\ to $z<0.7$, since at higher redshifts
it leads to duty cycles above unity.
With this model (inspired by that of Aird et al.\ 2011), we are able
to obtain a rising $U(M_{\rm BH})$ at $z=0.5$. However, the prediction
for the observed $P(\lambda)$\ disagrees with the data of K06, H09, and KH09,
a discrepancy also noted by Aird et al.\ (2011).
A final tension arises at the high end of the black hole mass
function, where the models tend to overpredict the data.
While models with a mass-independent $P(\lambda)$ are acceptable
within the estimated
observational uncertainties, adding mass dependence of
$\lambda_c$ to produce a rising $U(M_{\rm BH})$ leads to a
substantial overprediction of the abundance of the most massive black
holes. We have mitigated this problem in
our G$(z,M_{\rm BH})$\ model by also increasing the radiative efficiency
at high $M_{\rm BH}$\ (Fig.~\ref{fig|BHMFcomparison}).
This solution has some observational support, as discussed
in Section~\ref{subsec|PLmassdependence}. On the other hand, including mergers as
calculated in Section~\ref{subsec|Mergers} worsens the overprediction
of $\Phi(M_{\rm BH})$ at high masses (Fig.~\ref{fig|Mergers}).
There are other factors that may ameliorate the discrepancy between
models and observational determinations of the abundance of the most
massive black holes.
A $\lambda$-dependent (or simply lower) bolometric correction may resolve this
tension, as discussed in Section~\ref{subsec|kbolLambda}.
The model shown by the dotted curve in Fig.~\ref{fig|NMKbolLambda},
which includes mergers, is acceptable within current uncertainties.
The high end of the mass function relies on
extrapolation of the $M_{\rm BH}$-bulge relations into a regime where
there are few calibrating galaxies
(Schulze \& Gebhardt 2011).
Another effect that can help fitting the AGN luminosity function
with a steeper decline of the black hole mass function at high masses
is anisotropic emission by AGNs, which would give rise to apparent
values of $\lambda$ (i.e., inferred from the observed flux in our
direction under the assumption of isotropy) that are occasionally larger
than unity. The majority of the most luminous AGN would then correspond
to objects that have their brightest direction of emission pointing to
us, rather than to AGN with the highest black hole masses,
and the mass accretion rates of these AGN would be lower than
their observed luminosities suggest.
(Note that in
our models we have imposed a cutoff on the $P(\lambda)$ distribution at
$\lambda > 1$.)
If, in light of this discussion, we take a rather generous
view of the systematic uncertainties in the observational constraints
we have considered, then our G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$\ reference models may
be viewed as at least moderately successful, while the other
reference models --- $\delta$, G, G+P, and G$(z)$\ --- all fail
drastically on at least one observable. Table~\ref{table|models2}
summarizes the observational comparison based on an admittedly
subjective assessment of the results shown in
Figures~\ref{fig|BHMFandDuty}, \ref{fig|CompareKollmeier1}, \ref{fig|CompareKauff1}, \ref{fig|DutyCyclez0}, \ref{fig|BHMFcomparison}, and \ref{fig|CompareKollKauff}. We consider as constraints
the SWM estimate of the local black hole mass function, the
high-$z$ Eddington ratio distributions from K06, the low-$z$
Eddington ratio distributions from H09 and KH09, and the
duty cycle estimates from active galaxy fractions at $z\approx 0.2$
shown in Figure~\ref{fig|DutyCyclez0}. Recall that all models reproduce our input
AGN luminosity function by construction. We assign a
$\checkmark$ when a model reasonably describes an observation
with some allowance for systematic uncertainty, an X when it
clearly fails, and a $-$ for intermediate cases.
The non-evolving models and G$(z)$\ model all fail to match
the low-$z$ $P(\lambda)$\ or the low-$z$ $U(M_{\rm BH},L_{\rm min})$,
or both. The G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$\ models have no such drastic
failures, though they are far from perfect matches to the data.
However, none of our models reproduce the trend of duty cycle
with black hole mass illustrated in Figure~\ref{fig|DutyCycleHighz}, though the
G$(z,M_{\rm BH})$\ model is the least discrepant.
\begin{table*}
\begin{tabular}{|l|l|l|l|l}
\hline
Model & $\Phi_{\rm BH}(M_{\rm BH},z)$ $, \, z=0$ & $P(\lambda|M_{\rm BH},z)$ $, \, z>0.5$ & $P(\lambda|M_{\rm BH},z)$ $, \, z\le 0.5$ & $U(M_{\rm BH},L_{\rm min},z)$ \\
\hline
\hline
$\delta$ & $\checkmark$ & X & X & X \\
G & $\checkmark$ & $\checkmark$ & X & X \\
G+P & $\checkmark$ & $\checkmark$ & X & X \\
G(z) & -- & $\checkmark$ & X & $\checkmark$ \\
G(z,$M_{\rm BH}$) & $\checkmark$ & -- & -- & $\checkmark$ \\
G+P(z,$M_{\rm BH}$) & $\checkmark$ & $\checkmark$ & $\checkmark$ & -- \\
\hline
\end{tabular}
\caption{List of the Reference Models listed in Table~\ref{table|models} along
with a qualitative assessment of their agreement with the data. We assign a
$\checkmark$ when a model reasonably describes an observation
with some allowance for systematic uncertainty, an X when it
clearly fails, and a $-$ for intermediate cases. The $P(\lambda|M_{\rm BH},z)$\ at $z>0.5$ column refers to the K06 data (Figures~\ref{fig|CompareKollmeier1} and \ref{fig|CompareKollKauff}). The $P(\lambda|M_{\rm BH},z)$\ at $z\le 0.5$ refers instead to the H09 and KH09 data (Figures~\ref{fig|CompareKauff1} and
\ref{fig|CompareKollKauff}). Finally, the $U(M_{\rm BH},L_{\rm min},z)$ column refers to the multiple data sets reported in Figures~\ref{fig|DutyCyclez0} and \ref{fig|DutyCycleHighz}.}
\label{table|models2}
\end{table*}
\section{CONCLUSIONS}
\label{sec|conclu}
We have extended the formalism of continuity-equation modeling
of the black hole and AGN populations to allow a distribution
of Eddington ratios $P(\lambda)$. With this broader class of models
we have addressed two new categories of observations, direct
estimates of Eddington ratio distributions of active black holes
and estimates of duty cycles from active galaxy fractions.
Both of these categories have been areas of intense observational
investigation over the last five years, and as representative
examples we have concentrated on $P(\lambda)$\ estimates from
Kollmeier et al.\ (2006; K06) at $z \geq 1$,
Hickox et al.\ (2009; H09) at $z \approx 0.5$, and
Kauffmann \& Heckman (2009; KH09) at $z \approx 1$, and on
active galaxy fraction data from Bundy et al.\ (2008) and
Xue et al.\ (2010) at $z \geq 0.5$ and from
Goulding et al.\ (2009), Kauffmann et al.\ (2003), and
Schulze \& Wisotzki (2010) at low redshift.
We account for the effective luminosity thresholds of these
analyses in our model predictions, though the thresholds are
not always clearly defined.
As forms for $P(\lambda)$\ we consider a Gaussian in $\log\lambda$ (G)
and a Gaussian with a power-law extension to low-$\lambda$ (G+P).
If the radiative efficiency $\epsilon$ and characteristic
Eddington ratio $\lambda_c$ (where the Gaussian peaks) are
held fixed, then changing from a single $\lambda$ value to either
of these distributions has little impact on the evolution of
the black hole mass function inferred from continuity-equation modeling.
As in the single-$\lambda$ models of SWM (and similar models by
Shankar et al.\ 2004 and Marconi et al.\ 2004), we find that models
incorporating the observed AGN luminosity function and parameter
values $\epsilon \approx 0.07$ and $\lambda_c \approx 0.25$
yield a good match to observational estimates of the black hole
mass function in the local Universe.
The predicted black hole duty cycle declines rapidly with
decreasing redshift at $z<2$, and at redshifts $z \leq 1$
it declines sharply with increasing black hole mass over
the range $10^7 M_\odot - 10^9 M_\odot$
(``downsizing'' evolution).
An unattractive feature of the G+P model is that it requires
a high space density of massive black holes already present
at $z=6$, since only the small fraction of black holes with
high $\lambda$ are luminous enough to contribute to the
observed luminosity range at this redshift. This large ``seed''
population appears physically unrealistic, and the low implied
duty cycle for high-luminosity quasars contradicts evidence from
their strong observed clustering at $z \approx 4$
(Shen et al.\ 2007; White et al.\ 2008; Shankar et al.\ 2010b).
We conclude (in agreement with Cao 2010) that any extended low-$\lambda$
tail of $P(\lambda)$\ must develop at lower redshifts ($z \leq 3$) rather than
being a redshift-independent feature of black hole fueling processes.
Our models with redshift-independent $P(\lambda)$\ predict duty cycles at
$z\leq 1$ that are far below observational estimates from active
galaxy fractions, and they do not match the low-$z$ Eddington
ratio distributions estimated by H09 and KH09.
Motivated by these discrepancies, we introduce the G$(z)$\ model with
a redshift-dependent $\lambda_c$ (Eq.~\ref{eq|PLz}), shifting
$P(\lambda)$\ to lower Eddington ratios at low redshift.
This model yields better agreement with the observations.
The low-redshift $P(\lambda)$\ remains narrow compared to the H09 and KH09
estimates, and the model still predicts a duty cycle
that declines with increasing black hole mass, in contradiction
to the Bundy et al.\ (2008) and Xue et al.\ (2010) finding
of higher AGN fractions in more massive galaxies.
We therefore introduce an additional {\it mass} dependence of
$\lambda_c$ (Eq.~\ref{eq|PLmassDependent};
$\lambda_c \propto M_{\rm BH}^{-0.3}$) at each redshift,
which, relative to mass-independent models, maps massive black
holes to lower luminosity, more numerous AGN. At the same time,
we introduce a steady broadening of the Gaussian $P(\lambda)$\ towards
low redshift (Eq.~\ref{eq|Sigma}), or, for the G+P model,
a steady drop in the minimum Eddington ratio (Eq.~\ref{eq|LambdaMin})
that removes the need for an unrealistic seed population at $z=6$.
The predictions of the resulting models, G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$,
are fairly similar. Both models achieve a reasonable match to
the H09 and KH09 Eddington ratio distributions, though they
underpredict the KH09 distribution at $\lambda \leq 10^{-2.5}$.
They produce approximate agreement with the low-redshift duty
cycle estimates if we adopt the large (factors of $10-20$) ratios
of Type II to Type I AGN advocated by Greene \& Ho (2009),
though there are still factor of several discrepancies for
some data sets at some $M_{\rm BH}$ values.
Both models predict duty cycles that are flat or weakly
rising with black hole mass at $z>1$, thus improving the
agreement with Bundy et al.\ (2008) and Xue et al.\ (2010),
though neither model can reproduce the Xue et al.\ (2010)
trend at $z=0.5$.
Our models with mass-dependent $P(\lambda|M_{\rm BH},z)$\ exhibit tension
with the K06 Eddington ratio distributions,
especially at $z\approx 1$, since the redshift and mass
dependence of $\lambda_c$ drive the predicted $P(\lambda)$\ distributions
to peak at $\log\lambda \approx -1$ while the
observed distributions peak at $\log\lambda \approx -0.6$.
Adding a steep ($P \propto \lambda^{-0.9}$) power-law at late times,
similar to the model of Aird et al.\ (2011), can better
match the Xue et al.\ (2010) duty cycle data at $z=0.5$, but it also spoils
the match with the H09 and KH09 Eddington ratio distributions.
Boosting the duty cycle, and thus the growth, of massive
black holes tends to overproduce the high-mass end of the
local mass function relative to observational estimates.
Our standard versions of the G$(z,M_{\rm BH})$\ and G+P$(z,M_{\rm BH})$\ models therefore
incorporate a mass-dependent radiative efficiency
(Eq.~\ref{eq|radefficiencyMass}), following Cao \& Li (2008).
The higher efficiency assumed for higher mass black holes
reduces their inferred growth, restoring agreement with the
local mass function. Alternatively, a $\lambda$-dependent
bolometric correction (Vasudevan \& Fabian 2007) can lower the
inferred growth of massive (hence lower $\lambda$) black holes
in these models, obviating the need for mass-dependent $\epsilon$.
However, black hole mergers at the rate suggested by black hole
merger statistics also raise the high mass end of the mass
function at low redshifts, so in a complete model the mass-dependent
$\epsilon$ may be be required even with the $\lambda$-dependent
bolometric correction. Furthermore, Davis \& Laor (2011;
but see also Raimundo et al. 2011 and Laor \& Davis 2011) have
presented direct evidence for mass-dependent $\epsilon$ from
quasar spectral energy distributions, and such a dependence
could also explain the discrepancy between the $\epsilon \ga 0.2$
inferred for luminous, strongly clustered quasars at $z=4$
(Shankar et al.\ 2010b) and the average $\epsilon \approx 0.07$
implied by matching the local black hole mass density
(SWM and numerous references therein). Each of these arguments
for mass-dependent $\epsilon$ rests on somewhat shaky ground, but
they all point in the same direction.
At $z>1$, the black hole mass functions implied by our models
are fairly insensitive to mergers and only moderately sensitive
to other model assumptions.
In agreement with previous studies, we find that the growth of the
black hole population tracks the overall evolution of the cosmic
star formation rate. We also find generally good agreement
between the mean {\it specific} black hole growth rates
$\langle \dot{M}_{\rm BH}\rangle / M_{\rm BH}$ and
Karim et al.'s (2011) estimates of the specific star-formation
rates of star-forming galaxies over the full range $0.3 \leq z \leq 3$,
if we simply translate stellar masses to black hole masses
with a constant scaling factor of $10^{-3}$.
Over the last five years, measurements of AGN clustering have
improved dramatically in precision, redshift extent, and luminosity range.
These measurements provide valuable constraints on the relation between
active black holes and their host dark matter halos, which we have
previously explored in the context of single-$\lambda$ accretion
models (Shankar et al.\ 2010b,c). We will extend our clustering
studies to include $P(\lambda)$\ distributions and mergers in future work.
With $P(\lambda)$\ and radiative efficiency allowed to depend on redshift
and black hole mass, our models have become rather elaborate
despite their simple physical basis. This complexity reflects
the growing richness of the observational data, especially the
measurements of Eddington ratio distributions and active galaxy
fractions over a wide range of redshift, mass, and luminosity,
from a variety of data sets. In order of decreasing robustness,
our key qualitative conclusions are
(a) that the characteristic Eddington ratio $\lambda_c$ declines
at low redshift, (b) that the $P(\lambda)$\ distribution broadens
at low redshift, (c) that more massive black holes have lower $\lambda_c$,
and (d) that more massive black holes have higher radiative efficiency.
Despite the flexibility of our
models, and despite investigating many variants beyond those
discussed in the paper, we have not found a model that fully reproduces
all of the observational constraints we have considered.
The remaining discrepancies presumably reflect some combination
of inadequate models and systematic errors in the data
sets, as discussed in detail in Section~\ref{sec|discu}.
On the model side, we have assumed restricted functional
forms for the mass and redshift dependence of $P(\lambda)$, and more
general behavior or sharper evolutionary transitions may
be required to match the data. Observationally, estimates
of black hole masses and bolometric luminosities are both
subject to systematic uncertainties, and even when these
estimates are correct in the mean, scatter can have important
effects on inferred trends and distributions. Many of the tensions
between the models and the data and among the data sets
themselves revolve around the seemingly disparate trends found
for optically luminous, broad-line quasars and varieties
of low luminosity, Type II AGN. We have followed standard
practice in relating these two populations by a simple
obscuration factor, but a more nuanced relation between the
different categories of active black holes (e.g., Trump et al.\ 2011)
may be crucial to resolving some of the tensions highlighted here.
Continuity-equation models draw on the inevitable link between luminosity
and mass accretion to tie the observable population of AGN to the
evolving population of supermassive black holes that power them.
They provide a powerful framework for linking empirical studies that
probe a variety of observables across a wide span of redshift,
luminosity, and black hole mass. As these empirical studies continue
to improve in precision, dynamic range, and control of systematic
uncertainties, they will refine the models
into a tightly constrained history of the cosmic black hole population.
\section*{acknowledgments}
FS acknowledges support from the Alexander von Humboldt Foundation and a Marie Curie Grant.
We also acknowledge support from NASA Grant NNG05GH77G and NSF grant
AST-1009505. DW acknowledges support of an AMIAS membership at the
Institute for Advanced Study during part of this work.
We thank Elisa Binotto, Adam Steed, Jaiyul Yoo,
Yue Shen, Chris Onken, Zheng Zheng, Jeremy Tinker, Pavel Denisenkov, Roderik Overzier,
Zoltan Haiman, Massimo Dotti, Peter Behroozi, Lucia Ballo, Ana Babi\'{c}, Lance Miller,
Vincenzo Mainieri for many interesting and helpful discussions.
We thank the anonymous referee for a constructive report
that helped us improve the clarity of the paper.
|
\section{introduction}
Though the notion of a categorical model of dependent type theory was known for
quite some time now, it is only in recent years that it was realized that the
extra categorical structure required to model the structure of equality in
dependent type theory corresponds to the structure of weak factorization
equivalence, occurring in Quillen's model categories (\cite[p.2]{Ob}). This
connection is the basis for V. Voevodsky project known as \emph{univalent
foundations} whose main objective is to give a foundation of mathematics based
on dependent type theory, which is intrinsically homotopical, in which types
are interpreted not as sets, but rather as homotopy types (cf.). A central
ideal in Voevodsky's univalent foundations is the extension of Martin-L\"of's
dependent type theory by a ``homotopy theory reflection principle'', known as
the \emph{Univalence Axiom}. Roughly speaking, the Univalence Axiom is the
condition that the identity type between two types is naturally weakly
equivalent to the type of weak equivalences between these types (V. Voevodsky,
talk at UPENN, May 2011).
Within the category (or, rather, the model category) of simplicial sets
$sSets$, Voevodsky constructs a model of Martin-L\"of dependent type theory,
satisfying also the Univalence Axiom. The models constructed in this way are
called the standard univalent models (cf. Definition 3.2). During a
mini-workshop around these developments held in Oberwolfach in 2010 the
following question was raised: ``Does UA have models in other categories (e.g.,
1-topoi) not equivalent to the standard one?'', \cite[p.27]{Ob}. Though this
question is probably referring to a univalent universe (for type theory), it
seems to be meaningful also if taken literally. It turns out that the
Univalence Axiom can be given a precise meaning in the framework of Quillen's
model categories (provided they are locally Cartesian closed). It is then
meaningful to ask whether such a model category satisfies the Univalence Axiom.
There are two main parts to this note. In the first of these parts (Section
\ref{univ}) we give an interpretation of the notion of a univalent fibration in
a purely category theoretic language. To formulate this notion we introduce,
for a model category $\mathfrak C$, a correspondence $Hom^{(w)}(Z\times B,C): \mathfrak C\longrightarrow
Sets$, intended to capture the class of weak equivalences between given fibrant
objects $B,C\in \Ob\mathfrak C$. We then show that, given a fibration $p:C\xrightarrow{(f)} B$,
if $Hom^{(w)}_{B\times B}(-\times B\times C,C\times B)$ is a representable functor
(in the slice category $\mathfrak C/B\times B$), the ``obvious'' morphism (in $\mathfrak C/B\times B$) from
the diagonal $B_\delta$ to $Hom_{B\times B}(B\times C,C\times B)$ factors
``naturally'' (and uniquely in that sense) through the object representing this
functor, $((C\times B)^{B\times C})_w$. We can then define the fibration $p$ to
be \emph{univalent} if the morphism $m:B_{\delta}\longrightarrow ((C\times B)^{B\times
C})_w$ is a weak equivalence. This construction (or a closely related one) is
probably known to experts in the field, but since we could not find any
reference suitable for our purposes we give it in textbook detail.
We then introduce the notion of a locally (w/f)-Caretsian closed model
category, which is a locally Cartesian closed model category with the
additional property that $Hom^{(w)}_{B\times B}(-\times B\times C,C\times B)$ is a
representable functor for any fibrant objects $B,C$ and fibration $p:C\longrightarrow B$.
We observe that in a posetal (w/f)-Cartesian closed model category all
fibrations are univalent in the above sense. This is, of course, to be expected
in view of Voevodsky's informal description of a univalent fibration as
``...one of which every other fibration is a pullback in at most one way (up to
homotopy)''. Apparently, this should suffice to assure that any posetal
(w/f)-Cartesian closed model category satisfies the Univalence Axiom. But this
does not provide, to our taste, a satisfying analogy with Voevodsky's
construction. Such an analogy should have a natural interpretation of all the
key features in Voevodsky's construction. To our understanding one such feature
of univalent models is that they come equipped with a universal (univalent)
fibration, of which all ``small'' fibrations are a pullback (in a unique way),
\cite[Theorem 3.5]{Voev}. So our aim is to show, in addition, that such a
universal fibration exists in our model category (with respect to an
appropriate notion of smallness).
The second of the main parts of the paper (Section \ref{qtc}) is dedicated to a
self-contained construction of a posetal locally (w/f)-Cartesian closed model
category, $\QtN_c$. This construction is a special case of a more general
construction introduced in \cite{GaHa}. In \cite[p.8]{Ob} Voevodsky writes:
``Now for any $A, B : U$ , it is possible to construct a term $\theta : pathsU
(A, B) \longrightarrow weq(A, B)$... The Univalence Axiom states that the map $\theta$
should itself be a weak equivalence for every $A, B : U$''. The map $\theta$ in
Voevodsky's quote corresponds (to the best of our understanding) to the
morphism $m$ appearing in the factorisation of the ``obvious morphism''
mentioned above. It is now obvious that this formulation of the Univalence
Axiom is satisfied (in a rather trivial sense) in $\QtN_c$. To fulfill our goals,
it remains to construct an analogue in $\QtN_c$ of Voevodsky's universe of
``small'' fibration (those fibrations all of whose fibers are of cardinality
smaller than $\alpha$ for some cardinal $\alpha$). To that end we suggest a
(possibly over-simplified) notion of smallness for fibrations in a posetal
model category, and show that with this definition $\QtN_c$ admits a universe of
small fibration (which is automatically univalent).
Admittedly, the model category $\QtN_c$ may be too simple an object to be of real
interest. In \cite{GaHa} we suggest a construction of a (c)-(f)-(w)-labelled
category analogous to that of $\QtN_c$ resulting in a non-posetal category whose
slices are equivalent to those of $\QtN_c$. This category satisfies axioms
(M1)-(M5) of Quillen's model categories, but does not have products (and
co-products). We ask whether this richer category can be embedded in a model
category, and whether such a model category would satisfy the Univalence Axiom,
as formulated in this note.
It should be made clear that none of the authors of this note is familiar with
type theory and its categorical models. When we realized, moreover, that
formally accurate literature on Voevodsky's univalent foundations exists only
in the form of Coq code, we decided to base our homotopy theoretic
interpretation of the Univalence Axiom on the somewhat less formal presentation
appearing, e.g., in \cite{Voev}, \cite{Ob} and similar sources whose language
is closer to the categorical language for which we were aiming. To compensate
for the lack of precise references, we have taken some pains to give a detailed
formal account of our interpretation of those sources. M. Warren's comments and
clarifications, \cite{Warren}, were of great help to us, but all mistakes, are
- of course - ours.
A couple of words concerning terminology and notation are in place. In this
text we refer to Quillen's axiomatization of model categories, as it appears in
\cite{Qui}. Our usage of ``Axiom (M0)$\dots$(M5)'' refers to Quillen's
enumeration of his axioms in that book. Our commutative diagram notation is
pretty standard, and is explained in detail in \cite{GaHa}. The labeling of
arrows, (c) for co-fibrations, (f) for fibrations and (w) for weak
equivalences, is borrowed from N. Durov.
\section{Cartesian closed posetal categories}\label{CCC}
Given a category $\mathfrak C$ and $B,C\in \Ob\mathfrak C$, it is often desirable to treat
$Hom(B,C)$ as an object of the category: this is a natural requirement, as it
is inconvenient, while working in $\mathfrak C$, to be constantly required to work with
elements
external to $\mathfrak C$, namely, working with $Hom$-sets merely as sets. A category
$\mathfrak C$ is Cartesian closed if it is closed under ``exponentiation'', namely,
that given $B,C\in \Ob \mathfrak C$, an object $C^B$ (satisfying certain category
theoretic properties to be explained shortly) exists. As notation suggests, the
object $C^B$ is supposed to represent the set $Hom(B,C)$. We will now explain
this in more detail:
A category $\mathfrak C$ (with finite limits, or at least binary products) is called
{\em Cartesian closed} if
for every $B,C\in \Ob\mathfrak C$ there exist an object, denoted $C^B$,
and an arrow $\epsilon:C^B \times B \longrightarrow C$ such that for every
object $D$ and arrow $D\times B \xrightarrow{g} C$, there is a unique arrow
$f:D\longrightarrow C^B$
such that $D\times B\xrightarrow {f \times \id_B} C^B \times B \xrightarrow
{\,\epsilon\,} C$
and $D\times B \xrightarrow {\,g\,} C$ coincide \cite[6.2,p.108]{Awodey}.
\begin{figure}[H]
\centerline{
\xymatrix @R=5pc @C=5pc{
B \ar[r]^{\id_B} & B \\
D\times B \ar@/_2pc/|-(0.3)g[rr] \ar@{.>}[r] \ar[u]\ar[d] & C^B\times B
\ar[r]|-{\epsilon} \ar[u]\ar[d]& C \\
D \ar@{.>}[r]^f & C^B
}}
\caption{The existence of the arrow $D\xrightarrow{f} C^B$ assures the
existence of the arrow $D\times B\xrightarrow{f\times \id_B} C^B\times B$ by
the universal property of $C^B\times B$.}\label{CC}
\end{figure}
Given $B,C\in \mathfrak C$, Figure \ref{CC} implies the existence of a bijective
correspondence, functorial in $D$, between $Hom(D\times B,C)$ and $Hom(D,C^B)$,
which we can write as:
\begin{equation*}\tag{$*$}
Hom(D\times B,C)\equiv Hom(D,C^B)
\end{equation*}
In the above equation we say that $C^B$ is an \emph{object representing} the
functor $Hom(-\times B,C):\mathfrak C\longrightarrow Sets$, i.e. that this functor is naturally
equivalent to the functor $Hom(-,C^B)$. This equation is of importance not only
in understanding the ideology behind the definition of a Cartesian closed
category, but will also play an important role in our interpretation of the
Univalence Axiom in a model category.
To see why $C^B$ can be, in many cases, identified with the set $Hom(B,C)$
consider, in the equation $(*)$ the terminal object, $\top$ (for the variable
$D$). We get an equivalence of categories:
\[
Hom(B,C)\equiv Hom(\top\times B, C) \equiv Hom(\top, C^B)
\]
which, in many cases (e.g., in the category $Sets$, or in the category $Top$
where $\top$ is a point) gives:
\[
Hom(B,C)=Hom(\top,C^B)=C^B
\]
This explains the general category theoretic convention of identifying
exponents with $Hom$-sets.
In this note we will be interested, mainly, in posetal model categories. We
conclude this section with a discussion of a posetal category being Cartesian
closed. Recall that a category $\mathfrak C$ is \emph{posetal} if arrows are unique
whenever they exist. Namely, given $B,C\in \Ob\mathfrak C$ there exists at most one
$f\in \Mor\mathfrak C$ such that $B\xrightarrow{f} C$. Thus, in a posetal category all
diagrams are commutative, and therefore, as can be seen in Figure \ref{CC}, if
$\mathfrak C$ is posetal, to verify that $\mathfrak C$ is Cartesian closed it is enough to
verify that for any $B,C\in \Ob \mathfrak C$
there exists an object $C^B$ such that $C^B \times B \longrightarrow C$ and such that for
every
object $D$, $D\times B \longrightarrow C$ implies $D\longrightarrow C^B$.
Given a category $\mathfrak C$ and $A\in \Ob\mathfrak C$, the \emph{slice of $\mathfrak C$ over $A$},
denoted $\mathfrak C/A$ is the category of arrows $B\longrightarrow A$: its objects are arrows
$B\longrightarrow A$ in $\mathfrak C$ and an arrow from $B\longrightarrow A$ to $C\longrightarrow A$ is an arrow in
$\mathfrak C$ making the triangular diagram commute. For a posetal category the slice
$\mathfrak C/A$ can be identified with the full sub-category whose objects are all
$B\in \Ob\mathfrak C$ such that $B\longrightarrow A$.
A category is {\em locally Cartesian closed} if $\mathfrak C/A$ is
Cartesian closed for all $A\in \Ob\mathfrak C$, \cite[Prop.9.20,p.206]{Awodey}. Observe
that a posetal category with a terminal object is Cartesian closed if and only
if it is locally Cartesian closed. Indeed, $\mathfrak C$ has a terminal object $\top$
and $\mathfrak C/\top$ --- which is, by assumption, Cartesian closed - is merely $\mathfrak C$,
so locally Cartesian closed implies Cartesian closed. In the other direction,
if $\mathfrak C$ is Cartesian closed and $A\in \Ob\mathfrak C$ is any object, $B,C\in \Ob\mathfrak C/A$
then $C^B\times A\longrightarrow A$. Thus, $C^B\times A\in \Ob\mathfrak C/A$. So it remains to
verify that for any $D\in \Ob\mathfrak C/A$, if there exist an arrow $D\times B\longrightarrow C$
then there exist an arrow $D\longrightarrow C^B\times A$. By definition, there is an arrow
$D\longrightarrow C^B$, and since $D\in \Ob\mathfrak C/A$ there is an arrow $D\longrightarrow A$. By the
universal property of $C^B\times A$ this means that there is an arrow $D\longrightarrow
C^B\times A$, as required.
\section {The Univalence Axiom}\label{univ}
As explained in the introduction, the original formulation of the Univalence
Axiom is given in the language of type theory (and, apparently, its precise
formulation exists only in Coq code). The axiom asserts that, given a universe
of type theory, the homotopy theory of the types in this universe should be
fully and faithfully reflected by the equality on the universe. To prove that
the universes of type theory he constructs in the category $sSets$ are
univalent, Voevodsky proves, \cite[Theorem 3.5]{Voev}, that there is a
fibration universal for the class of \emph{small} fibrations, and that this
fibration is \emph{univalent}. Apparently, this statement is the right
reformulation of the Univalence Axiom in the context of the model category of
simplicial sets.
In order to generalize the Univalence Axiom to arbitrary (locally Cartesian
closed) model categories, one has to explain what it means for a fibration to
be univalent, and to define a suitable notion of smallness. Our first step is
to define (and explain) what is a univalent fibration in an arbitrary locally
Cartesian closed model category. We then show using this definition, that if
our locally Cartesian closed model category, $\mathfrak C$, is posetal, then every
fibration is univalent. Thus, to show that such a model category $\mathfrak C$ meets
the Univalence Axiom (for a suitable notion of smallness) it remains to show
that a universal fibration for all small fibrations exists. This section is
concluded with the observation that this is indeed the case for a natural
(though somewhat trivial) notion of smallness, provided $\mathfrak C$ is posetal.
\subsection{A model category object for weak equivalences}
Recall that Voevodsky's formulation of the Univalence Axiom takes place in the
category of simplicial sets. In order to reformulate this axiom in the more
general setting of model categories we have to set up a dictionary between
Voevodsky's terminology and the common terminology of model categories.
Apparently, such a translation is folklore to the experts, but since we were
unable to find a precise formulation meeting the level of generality need for
this note, we give the details. The main difficulty in this translation is the
definition of a univalent fibration. Since there is no literature on the
subject, our translation of this notion relies almost entirely on Voevodsky's
notes, \cite{Voev}, and some clarifications corresponded to us by Warren,
\cite{Warren}.
Let us recall Voevodsky's definition of a univalent fibration in the category
$sSets$ of simplicial sets, \cite[p.7]{Voev}:
\begin{quote}
For any morphism $q : E \rightarrow B$ consider the simplicial set
$\underline{Hom}_{B\times B} (E \times B, B \times E)$. If $q$ is a
fibration then it contains, as a union of connected components, a simplicial
subset
$weq(E\times B, B\times E)$ which corresponds to morphisms which are weak
equivalences.
The obvious morphism from the diagonal $ \delta δ : B \rightarrow B \times B$
to
$\underline{Hom}_{B\times B} (E \times B, B \times E)$ over $B \times B$
factors uniquely through a morphism
$m_q : B \rightarrow weq(E \times B, B \times E)$.
\end{quote}
In this terminology the fibration $q:E\longrightarrow B$ is \emph{univalent} if the
morphism $m_q:B\longrightarrow weq(E\times B,B\times E)$ is a weak equivalence (cf.
Definition 3.4 [ibid.])
Voevodsky's text translates readily into the language of Cartesian closed model
categories, with the possible exception of the definition of the object
$weq(E\times B,B\times E)$. In this subsection we perform this translation,
focusing on the model categorical definition of $weq(E\times B,B\times E)$. As
we will see, the object $weq(C,B)$ has much in common with the exponential
$C^B$, it is therefore convenient to introduce:
\begin{notation}
Given a model category $\mathfrak C$ and $B,C\in \Ob\mathfrak C$, the object $weq(C,B)$ will
be denoted $C^B_w$.
\end{notation}
For the sake of clarity, we explain the above text word for word. So let $\mathfrak C$
be a locally Cartesian closed model category, $E,B\in \Ob\mathfrak C$ and $q:E\longrightarrow B$ a
fibration. Let $E\times B$ be the product of $E$ and $B$ in $\mathfrak C$. This objects
comes with two morphisms: $E\times B\xrightarrow{\pr_E^{E\times B}} E$ and
$E\times B\xrightarrow{\pr_{B}^{E\times B}} B$.
Since a morphism into a product is uniquely determined by a pair of morphisms
into its components the following defines
a unique morphism:
$(q,id):E\times B \xrightarrow{q\circ \pr_E^{E\times B}\,\times\,
{\pr_{B}^{E\times B}}} B\times B$.
In set-theoretic notation, the morphism $q\circ \pr_E^{E\times B}\,\times\,
{\pr_{B}^{E\times B}}$ defined above is given by the mapping $(e,b)\mapsto
(q(e),b)$.
As an object of $\mathfrak C/B\times B$ this morphism is denoted by Voevodsky $E\times
B$.
In order to define the object $B\times E\in \Ob\mathfrak C/B\times B$ observe that (in
$\mathfrak C$) the object $B\times B$ comes equipped with two morphisms $\pr_1,\pr_2$
into each of its components. Thus, there is a morphism (in $\mathfrak C$)
$\tau_{B\times B}:B\times B \xrightarrow {\pr_2\times
\pr_1} B\times B$ (which can be thought of us the morphism permuting the
factors of the product). In set-theoretic notation $\tau_{B\times
B}(b_1,b_2)=(b_2,b_1)$.
The morphism $\tau\circ (q,id)$ as an object of $\mathfrak C/B\times B$ is denoted by
Voevodsky $B\times E$.
Thus, we have interpreted $B\times E$ and $E\times B$ as objects in the slice
category $\mathfrak C/B\times B$. In view of our discussion of exponentials in Section
\ref{CCC}, this allows us to identify $\underline{Hom}_{B\times B}(E\times
B,B\times E)$ with the object $((B\times E)^{E\times B})_{B\times B}$ (recall
that our assumption that $\mathfrak C$ is locally Cartesian closed assures that such an
object exists).
Recall that, in Voevodsky words, ``[the object $weq(E\times B, B\times E)$]
corresponds to morphisms [i.e., elements of $Hom_{B\times B}(E\times B, B\times
E)$]
which are weak equivalences [in the slice category ${\mathcal Qt}/B\times B$]''. Let us
now try to understand, in more generality, given a Cartesian closed model
category $\mathfrak C$ and objects $B,C\in \Ob\mathfrak C$ what should be the object $C^B_w$.
In the terminology used in Section \ref{CCC} Voevodsky's text should mean that
the object $C^B_w$ \emph{represents} the set of morphisms from $B$ to $C$
satisfying the additional requirement that these morphisms are weak
equivalences. In a Cartesian closed (model) category we identified $Hom(B,C)$
with the functor $Hom(-\times B,C)$. Since, in Voevodsky's text $C^B_w$ is a
sub-object of the exponential $C^B$, it is natural to try and identify $C^B_w$
with a sub-functor, let us denote it $Hom^{(w)}(-\times B,C)$, of $Hom(-\times
B,C)$. Moreover, any $Z\in \Ob\mathfrak C$ and morphism $f:B\longrightarrow C$ induces a morphism
$Z\times B\longrightarrow C$ given by $f\circ \pr_B^{Z\times B}$. It is, therefore,
reasonable to require that the same be true for the sub-functor
$Hom^{(w)}(-\times B,C)$. The ``obvious'' choice of letting $Hom^{(w)}(Z\times
B,C)$ be the set of all morphisms $h:Z\times B\xrightarrow{(w)} C$ does not have this
property. So the next best choice seems to be:
\begin{notation}
Given $Z,B,C\in \Ob\mathfrak C$, let
\[
Hom^{(w)}(Z\times B,C):=\{h:Z\times B\longrightarrow C\:|\: (\pr_Z^{Z\times B}\times
h):Z\times B\xrightarrow{(w)} Z\times C\}.
\]
\end{notation}
Observe that $Hom^{(w)}(Z\times B,C)\subseteq Hom(Z\times B,C)$ for all $Z$. We
do not know, however, whether --- in general --- it is functorial in $Z$. We
leave it as an exercise to the reader to show that if $\mathfrak C$ is right proper
(i.e., if the base change of a weak equivalence along a fibration is again a
weak equivalence), then $Hom(-\times B,C)$ is indeed functorial provided that
$B\xrightarrow{(f)} \top$ and $C\xrightarrow{(f)} \top$. We remind that $sSets$ is right proper,
and so is $Top$ --- and, more generally, any model category all of whose
objects are fibrant (see below) is right proper.
At all events, if $Hom^{(w)}(-\times B,C)$ is a functor, and as such it is
represented in $\mathfrak C$ we let $C^B_w$ denote the representing object. In
particular we obtain:
\begin{equation*}\tag{$**$}
Hom^{(w)}(Z\times B,C)\equiv Hom(Z,C^B_w)
\end{equation*}
\begin{definition}
Let $\mathfrak C$ a model category. Say that $\mathfrak C$ is (w/f)-Cartesian closed, if it is
Cartesian closed and, in addition, $Hom^{(w)}(-\times B,C):\mathfrak C\longrightarrow Sets$ is
represented (in the sense of $(**)$ above) for all fibrant $B,C\in \Ob \mathfrak C$
(i.e., the morphisms $B\xrightarrow{(f)} \top$ and $C\xrightarrow{(f)} \top$ into the terminal
object, $\top$ are fibrations). Say that $\mathfrak C$ is locally (w/f)-Cartesian
closed, if for any $X\in \Ob\mathfrak C$ the slice category $\mathfrak C/X$ is (w/f)-Cartesian
closed.
\end{definition}
\begin{rem}
The above definition is our straightforward interpretation of Voevosky's words
in the langauges of Cartesian closed model categories. This definition is
sufficeint for our purposes, as it is met by the model category $\QtN_c$,
constructed in the last section of this note. It is conceivable that, in the
general setting, a more accurate definition will be required.
\end{rem}
We will now show that if $\mathfrak C$ is a (locally) Cartesian closed model category
such that $Hom^{(w)}_{B\times B}(-\times (E\times B), B\times E)$ is
represented in $\mathfrak C/B\times B$ then the object representing this functor
satisfies the requirement in Voevodsky's text, namely, the ``obvious'' morphism
from the diagonal to $\underline{Hom}_{B\times B}(E\times B,B\times E)$ factors
uniquely through this representing object.
Before we proceed, some explanations are needed. Recall that we are working in
the slice category $\mathfrak C/B\times B$. Thus, the diagonal $\delta: B\to B\times B$ is an
object of $\mathfrak C/B\times B$, which we denote $B_{\delta}$. To avoid confusion, we
denote $E\underline\times B:=E\times B$ and $B\underline\times E:=B\times E$
(viewed as objects of $\mathfrak C/B\times B$, as explained above). We shall also let, given
an object $X\in \Ob\mathfrak C/B\times B$, $X_s\in \Ob\mathfrak C$ denote the source object of the
morphism (in $\mathfrak C$) corresponding to $X$. We have already explained in Section
\ref{CCC} in what sense $\underline{Hom}_{B\times B}(E\underline\times
B,B\underline\times E)$ can be viewed as an object of $\mathfrak C/B\times B$. So in order to
make Voevodsky's statement clear we only have to explain what is the ``obvious
morphism'' from the diagonal to $\underline{Hom}_{B\times B}(E\underline\times
B,B\underline\times E)$.
Consider the product $B_\delta \times_{B\times B} E\underline\times B$,
by definition it is the pullback of the morphisms
$B \xrightarrow \delta B\times B$ and $E\times B \xrightarrow{(q,id)} B\times
B$.
\begin{figure}[H]
\centerline{
\xymatrix @R=2pc @C=2pc{
(B_\delta \times_{B\times B} E\underline\times B)_s \ar[r]_(0.6){\pr_2}
\ar[d]^{\pr_1} & E\times B \ar[d]^{(q,\id_B)} \\
B\ar[r]^{\delta} \ar@/_2pc/[rr]|-\delta & B\times B\ar[r]^{\tau} & B\times B}}
\caption{}\label{2mor}
\end{figure}
It follows immediately from the fact that $\delta$ is the diagonal morphism and
from the definition of $\tau$ that $\tau\circ \delta\circ \pr_1=\delta\circ
\pr_1$. The right hand side morphism in the above equality corresponds in
$\mathfrak C/B\times B$ to the object $B_\delta\times (E\underline\times B)$, while the
composition $\tau\circ (q,\id_B)$ corresponds, by definition, to the object
$B\underline\times E$. The commutativity of the diagram of Figure \ref{2mor}
implied by the above equality means, by definition of $\mathfrak C/B\times B$ that the
morphism $\pr_2$ corresponds in $\mathfrak C/B\times B$ to a morphism $h:B_\delta\times
(E\underline\times B)\longrightarrow B\underline\times E$ in $\mathfrak C/B\times B$. By $(*)$ the
morphism $h$ corresponds to a morphism $\bar m_q:B_\delta \longrightarrow
\underline{Hom}_{B\times B} (E\underline\times B, B\underline\times E)$.
The morphism $\bar m_q$ is {\em the obvious morphism from the diagonal $\delta
δ : B \rightarrow B \times B$ to
$\underline{Hom}_{B\times B} (E\underline\times B, B\underline\times E)$ over
$B \times B$}.
Let us denote $\pi_1$ and $\pi_2$ the morphisms in $\mathfrak C/B\times B$ corresponding to
the morphisms $\pr_1$ and $\pr_2$ respectively (see Figure \ref{2mor}). Note
that these morphisms can be identified with the two canonical morphisms
associated to $B_{\delta}\times (E\underline\times B)$ as the pullback of
$B_\delta$ and $E\underline\times B$. Thus, we have a morphism (in $\mathfrak C/B\times B$)
$\pi_1\times(\tau\circ\pi_2):B_\delta\times (E\underline\times B) \longrightarrow
B_\delta\times (B\underline\times E)$. This morphism, by definition of
$\mathfrak C/B\times B$, arises from a morphism $\sigma: ((B_\delta\times (E\underline \times
B))_s\longrightarrow ((B_\delta \times (B\underline \times E))_s$ in $\mathfrak C$. This morphism
in $\mathfrak C$ is readily seen to be an isomorphism, and therefore a weak
equivalence. By the definition of the model structure on $\mathfrak C/B\times B$, a morphism
$X\longrightarrow Y$ in $\mathfrak C/B\times B$ is labelled (w), (f) or (c) if and only if the
corresponding morphism $X_s\longrightarrow Y_s$ in $\mathfrak C$ is labelled (w), (f) or (c)
respectively. Thus $\sigma$ (or, rather, $(\pi_1,\tau\circ \pi_2)$) is a weak
equivalence also in $\mathfrak C/B\times B$. By definition this means precisely that $h\in
Hom^{(w)}(B_\delta \times (E\underline\times B),B\underline\times E)$.
Therefore, $(**)$ implies that $h$ corresponds to a unique morphism
$m_q:B_\delta \longrightarrow ((E\underline\times B)^{(B\underline\times E)})_w$.
Finally, observe that in any (w/f)-Cartesian closed model category, $\mathfrak C$, and
for all fibrant $B,C\in \Ob\mathfrak C$ setting $D=C^B_w$ in $(**)$, we get that the
identity $\id: C^B_w\longrightarrow C^B_w$ is an element of $Hom(C^B_w,C^B_w)$ and
therefore also of $Hom^{(w)}(C^B_w\times B,C)$. Because $Hom^{(w)}(D\times
B,C)\subseteq Hom(D\times B,C)$ for all $D$, we get that $\id: C^B_w\longrightarrow C^B_w$
is an element of $Hom(C^B_w\times B,C)\equiv Hom(C^B_w,C^B)$. Thus, $\id:
C^B_w\longrightarrow C^B_w$ induces, through this last equivalence of functors, a morphism
$\id_*^{(B,C)} : C^B_w\longrightarrow C^B$
and a natural transformation $ \id_*^{(B,C)}:Hom^{(w)} (- \times B, C)\equiv
Hom(-,C^B_w)$ coinciding
with the natural transformation provided by the inclusion $Hom^{(w)}(- \times
B, C)\subseteq Hom(- \times B, C)$.
Combining this with the conclusion of the previous paragraph, we get that $\bar
m_q=\id_*^{(B,C)}\circ m_q$.
The identification of $\id: C^B_w\longrightarrow C^B_w$ as an element of $Hom(-\times
B,C)$ is natural in that sense. Requiring that $\bar m_q$ factors
\emph{naturally} through $C^B_w$ amounts, therefore, to the requirement that
this factorization is obtained via $\id_*^{(B,C)}$. We observe that with this
additional requirement this factorization is unique.
\subsection{The Univalence Axiom in posetal model categories}
Having defined the object $C^B_w$ for a locally (w/f)-Cartesian closed model
category $\mathfrak C$, we can define a fibration $p:E\longrightarrow B$ to be \emph{univalent} if
the morphism $\bar m_q:B_\delta\longrightarrow ((E\underline\times B)^{(B\underline\times
E)})_w$ is a weak equivalence. In this subsection we prove:
\begin{lem}
Let $\mathfrak C$ be a locally (w/f)-Cartesian closed posetal model category. Then
every fibration is univalent.
\end{lem}
\begin{proof}
First, observe that since $\mathfrak C$ is posetal for any object $B\in \Ob\mathfrak C$ the
product $B\times B$ is isomorphic to $B$. Indeed, by the universal property of
$B\times B$ there is a morphism $B\longrightarrow B\times B$, and since $B\times B\longrightarrow B$ we get that
$B\cong B\times B$ (because $\mathfrak C$ is posetal).
Recall that since $\mathfrak C$ is posetal, for any object $B$ the slice category
$\mathfrak C/B$ can be identified with the full subcategory whose objects are $\{A\in
\Ob\mathfrak C: A\longrightarrow B\}$. Namely, the morphism $A\longrightarrow B$, as an object in $\mathfrak C/B$ can
be identified with the object $A$ in $\mathfrak C$. In particular $B_\delta$ can be
identified with the object $B$, and given a fibration $E\xrightarrow{(f)} B$ the objects
$E\underline\times B$ and $B\underline\times E$ in $\mathfrak C/B\times B\equiv \mathfrak C/B$ are
isomorphic and both can be identified with the object $E\times B$ of $\mathfrak C$
(indeed, in a posetal model category $E\underline\times B$ and
$B\underline\times E$ are the same object since $(E\underline\times
B)_s=(B\underline\times E)_s$). So $\underline{Hom}_{B\times B}(E\underline\times
B,B\underline\times E)$ is isomorphic to the object $(E\times B)^{E\times B}$.
It will suffice to show that for any object $C\in \Ob\mathfrak C$ the exponent $C^C$ is
isomorphic to $\top$, the terminal object of $\mathfrak C$. Indeed, then
$\underline{Hom}_{B\times B}(B\underline\times E, E\underline\times B)=((E\times
B)^{(E\times B))}_{B\times B}=\top_{B\times B}$. But the terminal object of $\mathfrak C/B\times B$ is
$B\times B=B$. We get that the ``obvious morphism'' $h:B_\delta \longrightarrow
\underline{Hom}_{B\times B}(E\underline\times B,B\underline\times E)$ defined in the
previous subsection corresponds to the arrow $B\longrightarrow B$, so it is an
isomorphism, and therefore a weak equivalence. But in a posetal model category,
if $X\longrightarrow Y$ is an isomorphism then for all $Z$, if $X\longrightarrow Z\longrightarrow Y$ then $X\longrightarrow
Z$ and $Z\longrightarrow Y$ are both isomorphisms. In particular, the morphism $\bar
m_q:B_\delta \longrightarrow ((E\underline\times B)^{(B\underline\times E)})_w$ is an
isomorphism, and therefore a weak fibration.
It remains, therefore, to show that in a posetal Cartesian closed model
category $C^C\cong \top$ for all $C\in \mathfrak C$. Indeed, $\top\times C\cong C$,
implying $\top\times C\longrightarrow C$. So by Figure \ref{CCC} (with $D=\top$ and $B=C$)
we get an arrow $\top\longrightarrow C$, and $\mathfrak C$ being posetal we get $C^C\cong \top$.
\end{proof}
Having seen that in posetal locally Cartesian closed model categories the
notion of univalent fibrations degenerates, it remains to show that there
exists a fibration $p$ universal for the class of \emph{small} fibrations. Of
course, the notion of smallness in this context should be defined as well.
\begin{definition}
Let $\mathfrak C$ be a model category, Fix a morphism $\tilde U \xrightarrow{p} U$. A
morphism $Y \xrightarrow{f} X$ is $p$-small if $Y\xrightarrow{f} X$ fits in a
pull-back square:
\begin{figure}[H]
\centerline{
\xymatrix @R=2pc @C=2pc{
Y \ar@{.>}[r] \ar[d]|-f & \tilde U \ar[d]|-p \\
X\ar[r]|-{f_p} & U
}}
\caption{This is a pullback square, if for any morphisms $Z\longrightarrow X$ and $Z\longrightarrow
\tilde U$ making the diagram commute there is an arrow $Z\longrightarrow Y$ making the
diagram commute. }
\end{figure}
Say that $p$ is \emph{universal} (with respect to a pre-defined class of
\emph{small} fibrations) if the class of $p$-small fibrations contains all
small fibrations.
\end{definition}
Observe that in a posetal category, given morphisms $p$ and $f$ as in the above
definition, the morphism $X\xrightarrow{f_p} U$ is unique if it exists.
Therefore, $Y\xrightarrow{f} X$ is $p$-small if and only if $X\longrightarrow U$ and
$Y=\tilde U\times X$.
\begin{lem}\label{psmall}
Let ${\mathcal Qt}$ be a posetal model category. Consider the unique morphism $\varnothing\longrightarrow
\top$ and let $\tilde U$ be the unique object such that $\varnothing\xrightarrow{(wc)} \tilde U
\xrightarrow{(f)} \top$. Let $p$ denote the fibration $\tilde U \xrightarrow{(f)} \top$. Assume,
in addition, that all morphisms in ${\mathcal Qt}$ are co-fibrations. Then a fibration
$f:Y\longrightarrow X$ is $p$-small iff $\emptyset \xrightarrow{(wc)} Y$.
\end{lem}
\begin{proof}
The key to the proof is the following observation:
\noindent{\bf Claim} If $Z\longrightarrow \tilde U$ then $Z\xrightarrow{(wc)} \tilde U$.\\
\proof Let $Z\xrightarrow{(wc)} Z_{wc}\xrightarrow{(f)} \tilde U$ It will suffice to prove that
$\tilde U\longrightarrow Z$, since then $Z_{(wc)}$ is isomorphic to $\tilde U$ (member
that ${\mathcal Qt}$ is posetal). Indeed, consider the following diagram:
\begin{figure}[H]
\centerline{
\xymatrix @R=2pc @C=2pc{
\perp \ar[d]|-{(wc)} \ar[r] & Z_{wc} \ar[d]|-{(f)}\\
\tilde U \ar[r] \ar@{.>}[ur] & \tilde U
}}
\end{figure}
Finishing the proof of the claim \qed$_{\text {Claim}}$
Now, if $\perp\xrightarrow{(wc)} Y\xrightarrow{(f)} X$ (where $\perp$ is the initial object), then
$\perp\longrightarrow Y\rightthreetimes \tilde U \longrightarrow \top$, giving $Y\longrightarrow \tilde U$. So it suffices
to show that $Y=\tilde U\times X$. Because $\perp \xrightarrow{(wc)} Y \rightthreetimes \tilde U
\xrightarrow{(f)} \top$ we know that $Y\longrightarrow \tilde U$. So $Y\longrightarrow X\times \tilde U$. Let
$Y\xrightarrow{(wc)} Y_{wc} \xrightarrow{(f)} X\times \tilde U$. By the above claim $X\times \tilde
U \xrightarrow{(wc)} \tilde U$ and $Y_{wc}\xrightarrow{(wc)} \tilde U$. So by (M5):
\begin{figure}[H]
\centerline{
\xymatrix @R=2pc @C=2pc{
& Y_{wc} \ar[dl]|-{\therefore (w)} \ar[dr]|-{(wc)}\\
X\times \tilde U \ar[rr]|-{(wc)} & & \tilde U
}}
\caption{By (M5) the arrow $Y_{wc}\longrightarrow X\times \tilde U$ is a weak equivalence.}
\end{figure}
But, by assumption all arrows in ${\mathcal Qt}$ are co-fibrations, and we chose $Y_{wc}$
so that $Y_{wc}\xrightarrow{(f)} X\times \tilde U$. so $Y_{wc}\xrightarrow{(wcf)}X\times
\tilde U$, and since ${\mathcal Qt}$ is posetal, this implies that $Y_{wc}$ is isomorphic
to $X\times \tilde U$. We conclude that $Y\xrightarrow{(wc)} X\times \tilde U$.
Therefore $Y\longrightarrow X\times \tilde U \rightthreetimes Y\longrightarrow X$, giving an arrow $X\times
\tilde U \longrightarrow Y$, with the conclusion that $Y$ is isomorphic to the product, as
required.
In the other direction. If $Y\xrightarrow{(f)} X$ is $p$-small then $Y\longrightarrow \tilde U$, and
by the claim $Y\xrightarrow{(wc)} \tilde U$. Similarly, if $\perp \xrightarrow{(wc)} Y_{wc}\xrightarrow{(f)}
Y$ then $Y_{wc}\xrightarrow{(wc)} \tilde U$. So (M5), applied to the triangle $\tilde
U\longleftarrow Y_{wc}\longrightarrow Y\longrightarrow \tilde U$, assures that $Y_{wc}\xrightarrow{(w)} Y$.
Since, by assumption, all arrows are co-fibrations, we get
$Y_{wc}\xrightarrow{(wcf)} Y$, with the conclusion that $\perp\xrightarrow{(wc)} Y$, as
required.
\end{proof}
In order to conclude we have to give a reasonable notion of smallness ---
namely, to define when is a fibration $f:X\to Y$ in an arbitrary model category
\emph{small}. For reasons to be explained below we do not attempt to give a
definition of a small fibration in that generality. Rather, our goal is find
some (minimal) necessary conditions that such a class of fibrations should
satisfy. Since we are trying to interpret the Univalence Axiom, as it is
discussed in \cite{Voev}, it is natural that our analysis of the notion of
smallness be based on the definition of a universal fibration introduced there
(\cite{Voev}, p.6). In the category of simplicial sets a fibration $f:X\longrightarrow Y$
is \emph{small} (for some fixed cardinality $\alpha$) if all its fibers are of
cardinality smaller than $\alpha$.
Observe that (for most cardinalities) Voevodsky's definition of small
fibrations depends, e.g., on the choice of model of ZFC. This suggests that
there is no natural category theoretic counterpart exactly capturing this
definition. So, let us consider some obvious properties of Voevodsky's
definition:
\begin{enumerate}
\item The class of small fibrations is closed under finite products and
co-products.
\item Since co-fibrations are injective,
if
$f:X\longrightarrow Y$ is a small fibration, and $g:X'\longrightarrow X$ is a fibration and a
co-fibration then also $f\circ g: X'\longrightarrow Y$ is small.
\end{enumerate}
Thus, by the second point above, if ${\mathcal Qt}$ is a posetal model category all of whose
morphisms are co-fibrations, then for any small fibration $X\xrightarrow{(f)} Y$ if
$\perp\xrightarrow{(wc)} X_{wc} \xrightarrow{(f)} X$, then $X_{wc}\xrightarrow{(f)} Y$ should also be a small
fibration. Therefore, in any such model category, under any non-trivial
definition of small fibrations, some small fibrations will be of the form
$X_{wc}\xrightarrow{(f)} Y$ where $\perp \xrightarrow{(wc)} X_{wc}$. Moreover, in order to satisfy
the first of the above points we have to close the collection of trivial
co-fibrant objects, $X_{wc}$, such that there exists some small fibration $X_{wc}\xrightarrow{(f)}
Y$ under finite limits and co-limits. The properties of ${\mathcal Qt}$ assure that this
is still a collection of trivial co-fibrant objects (that the co-base change of
a weak co-fibration is a weak equivalence - and therefore in ${\mathcal Qt}$ a weak
co-fibration - follows from Axiom (M4) of model categories; that the product of
two trivial co-fibrant objects is a trivial co-fibrant objects is proved
precisely as in the claim of Lemma \ref{psmall}). Let $\mathcal S$ denote the
collection of trivial co-fibrant objects thus obtained. Consider
${\mathcal Qt}^{\mathcal S}$, the ``co-slice category over $\mathcal S$'', i.e.,
${\mathcal Qt}^{\mathcal S}$ is the full sub-category whose objects are all those $X\in
\Ob{\mathcal Qt}$ such that $S\xrightarrow{(f)} X$ for some $S\in \mathcal S$.
Then
${\mathcal Qt}^{\mathcal S}$ is still a model category (one only needs to check that
${\mathcal Qt}^{\mathcal S}$ is closed under finite limits and co-limits, which is
obvious). Moreover, as can be readily checked in Figure 1, since ${\mathcal Qt}$ is
Cartesian closed, so is ${\mathcal Qt}^{\mathcal S}$. Being posetal, ${\mathcal Qt}^{\mathcal S}$
is also locally Caretsian closed. In addition, if ${\mathcal Qt}$ is (locally)
(w/f)-Cartesian closed then, by putting $Z=C$ in $(**)$ we see that
${\mathcal Qt}^{\mathcal S}$ is also (locally) Cartesian closed.
It follows that, replacing ${\mathcal Qt}$ with ${\mathcal Qt}^{\mathcal S}$ we obtain a locally
Cartesian closed posetal model category all of whose trivial co-fibrant objects
are small, in the sense that whenever $\perp\xrightarrow{(wc)} X\xrightarrow{(f)} Y$ the fibration
$X\xrightarrow{(f)} Y$ is small. Of course, the model category ${\mathcal Qt}^{\mathcal S}$ may be
less interesting than the original category ${\mathcal Qt}$. But the above argument shows
that - at least for posetal model categories all of whose morphisms are
co-fibrations - it is possible to have a notion of smallness which corresponds
exactly to a fibration $X\xrightarrow{(f)} Y$ being small when $\perp\xrightarrow{(wc)} X$. Since,
as explained above, there cannot be a natural category theoretic definition of
smallness capturing precisely Voevodsky's notion of small fibration, we believe
that, given the level of generality we are working in, the above is as good an
approximation of this notion as could be expected.
We conclude that:
\begin{prp}\label{main}
Let ${\mathcal Qt}$ be a posetal model category all of whose morphisms are
co-fibrations. Let $\perp \xrightarrow{(wc)} \tilde U \xrightarrow{(f)} \top$, and define a
fibration $Y\xrightarrow{(f)} X$ to be small if $\perp \xrightarrow{(wc)} Y$. Then the fibration
$p:\tilde U\longrightarrow \top$ is universal. If, in addition, ${\mathcal Qt}$ is locally
(w/f)-Cartesian closed then ${\mathcal Qt}$ meets the Univalence Axiom, with respect to
the above notion of small fibrations.
\end{prp}
In the next section we give an example of a non-trivial model category
satisfying all the assumptions of Proposition \ref{main}.
\section{The model category $\QtN_c$}\label{qtc}
The main result of \cite{GaHa} is the construction of a (non-trivial) posetal
model category of classes of sets, $\textrm{QtNaamen}$. In this section we show that the
full sub-category, $\QtN_c$, of all co-fibrant objects meets all the assumptions
of Proposition \ref{main}. Indeed, this follows almost immediately from the
results of \cite{GaHa}, but for the sake of completeness, we give a
self-contained proof.
To simplify the exposition, and in order to avoid irrelevant foundational
issues, we give a slightly simplified version of the model category $\QtN_c$. Let
$\QtN_c$ be the category whose objects are the members of $\mathbb P( \mathbb P
(\mathbb N)):=\{X\subseteq \{M\subseteq \mathbb N\}\}$ and for $X,Y\in \Ob\QtN_c$
let $X\longrightarrow Y$ precisely when for every $x\in X$ there exists $y\in Y$ such that
$x\subseteq y$. We leave it as an easy exercise for the reader to verify that
this is indeed a (posetal) category.
\begin{claim}
The category $\QtN_c$ has limits. Direct limits are given by unions $X\vee Y =
X\cup Y$, and inverse limits are given by pointwise intersection, namely
$X\times Y = \{x\cap y: x\in X, y\in Y\}$. The same formulas hold for infinite
limits.
\end{claim}
\begin{proof}
This is straightforward. Assume, e.g. that we are given $X,Y$ and $Z\longrightarrow X$,
$Z\longrightarrow Y$. By definition, this means that for all $z\in Z$ there are $x\in X$,
$y\in Y$ such that $z\subseteq x$ and $z\subseteq y$. This means that for all
$z\in Z$ there are $x\in X$ and $y\in Y$ such that $z\subseteq x\cap y$. This
proves that $X\times Y$ as defined above is the inverse limit of $X$ and $Y$.
The proof for direct limits is similar.
\end{proof}
Now we endow $\QtN_c$ with a model structure. We do not attempt to justify the
intuition behind these definitions - this is done in some detail in
\cite{GaHa}. In order to meet the assumptions of Proposition \ref{main}, we
must require that all morphisms are labelled (c). So we now proceed to the (w)
and (f) labels.
For the definition of weak equivalences it is convenient to denote for $X, Y\in
\Ob \QtN_c$, $X\longrightarrow^* Y$ if for all $x\in X$ there exists $y\in Y$ such that
$|y\setminus x|<\aleph_0$. We now set $X\xrightarrow{(w)} Y$ if $X\longrightarrow Y$ and $Y\longrightarrow^*
X$. This definition obviously satisfies Axiom (M5) (2 out of 3). Also, if
$Z\xleftarrow{(wc)} X\longrightarrow Y$ then $Y\xrightarrow{(wc)} Z\vee Y$. Indeed, if $r\in Z\vee
Y$ then either $r\in Y$ in which case $r\setminus r=\varnothing$ or $r\in Z$, in which
case there is $x\in X$ such that $|r\setminus x|<\aleph_0$ but $X\longrightarrow Y$, so
there is $y\in Y$ such that $x\subseteq y$ and $|r\setminus y|\le |r\setminus
x|<\aleph_0$. This shows that the (wc)-part of Axiom (M4) is met by this
notation.
It remains to define the (f)-labelling: an arrow $X\longrightarrow Y$ is labelled (f) if
and only if for every $x\in X\cup\{\emptyset\}$, $y\in Y$ and a finite
subset $\{b_1,\dots ,b_n\}\subseteq y$ there exists $x'\in X$ such that
$(x\cap y)\cup\{b_1,...,b_n\}\subseteq x'$.
First, we observe:
\begin{claim}
If $X\xrightarrow{(wcf)} Y$ then $Y\longrightarrow X$.
\end{claim}
\begin{proof}
Let $y\in Y$. We have to show that there exists $x\in X$ such that $y\subseteq
x$. Let $x_0\in X$ be such that $z:=y\setminus x$ is finite, as provided by the
(w)-label. So the (f)-label, applied for $x_0,y$ and $z\subseteq y$ assures the
existence of $x$ with the desired property.
\end{proof}
This claim gives us, automatically, one part of (M1) - any arrow right-lifts
with respect to an isomorphism - one part of (M2) - any arrow $X\xrightarrow{(c)} Y$
decomposes as $X\xrightarrow{(c)} Y\xrightarrow{(wf)} Y$ and (M3) (it remains only to verify that
fibrations are stable under base-change). Axiom (M4) is also automatic. So we
are left with the $(wc)\rightthreetimes (f)$ part of (M1), the (wc)-(f) decomposition of
(M2) and the stability of fibrations under base change. All computations are
trivial, so we will be brief.
Let $X\xrightarrow{(wc)} Y$ and $W\xrightarrow{(f)} Z$ be such that $X\longrightarrow W$ and $Y\longrightarrow Z$. We
have to show that $Y\longrightarrow W$. So let $y\in Y$. Let $x\in X$ be such that
$b:=y\setminus x$ is finite. Let $w\in W$ be such that $x\subseteq w$. Let
$z\in Z$ be such that $y\subseteq z$. Apply the definition of (f)-arrows with
respect to $w,z$ and $b$. Then there exists $w'\in W$ such that $(w\cap z)\cup
b\subseteq w'$. So $y\subseteq w'$, as required. An essentially similar
argument shows that fibrations are stable under base-change.
To prove (M2), let $X\longrightarrow Y$ be any arrow. Let
\[
X_{wc}:=\{x\cup y_0: x\in X, (\exists y\in Y)(y_0\subseteq y), y_0\text{
finite}\}.
\]
Then $X\xrightarrow{(wc)} X_{wc}\xrightarrow{(f)} Y$, as can be readily checked.
We conclude that $\QtN_c$ is a posetal model category all of whose arrows are
co-fibrations. It is not trivial (in the sense that not all arrows are
fibrations) because $\varnothing\xrightarrow{(wc)} X$ is not a fibration unless $X=\varnothing$. Since it
is posetal, to show that it is locally Cartesian closed, it suffices to show
that it is Cartesian closed.
Define, for $C,B\in \Ob\QtN_c$:
\[
C^B:=\bigcup \{ Z: Z\times B \longrightarrow C ,\, A\longrightarrow Z\}.
\]
This is, obviously, an object in $\QtN_c$, so we need only check that
$C^B \times B \longrightarrow C$ and that for every
object $Z$, $Z\longrightarrow A$ and $Z\times B \longrightarrow C$ implies $Z\longrightarrow C^B$.
The latter is immediate by definition of $C^B$. The former
requires a little argument. We need to check that for every $d\in C^B$
and $b\in B$ there is a morphism $\{d\cap b\}\longrightarrow C$. By definition of $C^B$,
there exists $Z$ such that $Z\times B \longrightarrow C$ and $d\in Z$, i.e. $\{d\}\longrightarrow Z$.
By definition of the product $Z\times B$, this implies $\{d \cap b\} \longrightarrow C$,
as required.
\begin{rem}
Note that the above shows that $\QtN_c$ is, in particular, a logical model
category in the sense of \cite[Definition 23]{ArKa}. Consequently (cf. Theorem
26) $\QtN_c$ admits a sound interpretation of the syntax of type theory (though
the lack of non-trivial sections probably makes this interpretation trivial).
\end{rem}
All of the above shows that $\QtN_c$ is a posetal locally Cartesian closed model
category which is non-trivial (in the sense that not all morphisms are labelled
(fc)). So in order to apply Proposition \ref{main} it remains to show that it
is locally (w/f)-Cartesian closed. We prove:
\begin{claim}\label{claim5}
$ Z\times B\xrightarrow{(wc)} Z\times C$ iff for all $\{z\}\longrightarrow Z$, $\{z\}\times
B\xrightarrow{(wc)} \{z\}\times C$
\end{claim}
\begin{proof}
The right to left direction is immediate from the definition of (wc)-arrows, so
we prove the other direction. The arrow $Z\times B\xrightarrow{(wc)} Z\times C$ means
that:
\begin{itemize}
\item for any $z \in Z$, $b \in B$ exists $z' \in Z$ and $c'\in C$
such that $\{z \cap b\}\longrightarrow \{z'\cap c'\}$; and
\item for any $z''\in Z$, $c''\in C$ exists $z\in Z$, $b\in B$ such that
$ \{z''\cap c''\}\longrightarrow^* \{z \cap b\}$.
\end{itemize}
Observe that the first bullet (for fixed $z\in Z, b\in B$) gives $z \cap
b\subseteq z' \cap c'$, implying that $z \cap b\subseteq z \cap z'\cap
c'\subseteq z \cap c'$, therefore $\{z\}\times B \longrightarrow\{z\}\times C$.
Analogously, for fixed $z''\in Z, c''\in C$ the assumption $ \{z'' \cap
c''\}\longrightarrow^* \{z \cap b\}$
implies $\{z'' \cap c''\}\longrightarrow^* \{z'' \cap z \cap b\}\longrightarrow \{z''\cap b\}$.
Combining these two observations we get $\{z\}\times B \xrightarrow{(wc)} \{z\} \times
C$.
\end{proof}
Now, given $A\in \Ob\QtN_c$ and $B\longrightarrow A$, $C\longrightarrow A$, we define
\[
(C^B_w)/A = \bigcup \{Z : Z\times B \xrightarrow{(w)} Z\times C \}\times A
\]
and show that this is an object representing $Hom^{(w)}_A(-\times B,C)$ ($\QtN_c$
is trivially right proper, so this is indeed a functor). More precisely:
\begin{claim}
For all $Z\longrightarrow A$, we have
$Z \longrightarrow (C^B_w)/A$ if and only if $Z\times B \xrightarrow{(w)}
Z\times C.$
\end{claim}
\begin{proof}
The Right to left direction is immediate from the definition. So suppose $Z\longrightarrow
(C^B_w)/A$. We need to show that $Z\times B \xrightarrow{(w)} Z\times C$.
By Claim \ref{claim5}, this happens if for all $\{z\}\longrightarrow Z$,
$\{z\}\times B \xrightarrow{(w)} \{z\}\times C$.
But our assumption that $Z\longrightarrow (C^B_w)/A$ implies that $Z \longrightarrow Z'$ for some
$Z'$ such that $Z'\times B
\xrightarrow{(w)} Z'\times C$. So $\{z\}\longrightarrow Z'$, by Claim \ref{claim5} we are done.
\end{proof}
Combining everything together we get:
\begin{theorem}
There exists a non-trivial posetal model category satisfying the Univalence
Axiom.
\end{theorem}
Before continuing we remark that $\QtN_c$, as presented in this note is a full
sub-category of the category of co-fibrant objects in the model category $\textrm{QtNaamen}$
defined in \cite{GaHa}. The exact same proof of the above theorem would work
for the full category of co-fibrant objects in $\textrm{QtNaamen}$. Of course this category
captures the full homotopy structure of $\textrm{QtNaamen}$, and may - therefore - be a more
interesting example. We remark also that there does not seem to be anything
spcial about $\mathbb N$ or about $\aleph_0$ in the above construction (or in
the more genreal construction of $\textrm{QtNaamen}$). Apparently, the exact same
construction could be achieved for any regular cardinal $\lambda$ (in place of
$\aleph_0$) replacing, throughout ``finite'' by ``less than $\lambda$''. This
gives --- in addition to the formal discussion of smallness in the previous
sub-section --- an analogy with Voevodsky's notion of small fibrations: it is
not unreasonable (see the following paragraph) to think of the morphisms in the
resulting model category as a class of injections (satisfying certain
compatibility conditions), our definition of smallness implies that a fibration
is small precisely when every memebr of the class of these injections has a
domain smaller than $\lambda$.
To conclude, let us consider the category $\mathfrak C$, whose objects are $\Ob \QtN_c$
and such that $\Mor(X,Y)$ consists of the arrows $X\xrightarrow{\sigma} Y$ for
$X,Y\in \Ob\mathfrak C$ such that $\sigma:\bigcup X\longrightarrow \bigcup Y$ and $\sigma(X)\longrightarrow
Y$ is an arrow in $\Mor\QtN_c$ (where $\sigma(X):=\{\,\{\sigma(a):a\in x\}\,:\,x\in X\}$). The
category $\mathfrak C$ is, on the one hand, obvioulsy richer than $\QtN_c$ (it is not
posetal). But, on the other hand, it is readily seen that any slice of $\mathfrak C$ is
(naturally) equivalent to the corresponding slice of $\QtN_c$. This local model
structure induces naturally a (c)-(f)-(w) labeling on $\Mor(\mathfrak C)$ (see
\cite{GaHa} for the details) satsifying Quillen's axioms (M1)-(M5). But the
category $\mathfrak C$ does not have products and co-products. So we ask: \\
\noindent{\bf Question}: Is there a model category $\mathfrak C'$ such that the labeled
category $\mathfrak C$ described above embeds in $\mathfrak C'$? Does $\mathfrak C'$ satisfy the
Univalence Axiom? \\
\noindent\emph{Acknowledgement} We would like to thank N. Durov for his help in
the definition of (w/f)-Cartesian closed model categories.
\bibliographystyle{alpha}
|
\section{Introduction}
Among the 150 molecules that have been detected in the interstellar
medium (ISM) so far, a significant number are complex organic
molecules (here-after COMs), carbon bearing molecules with more than five
atoms. Several COMs have been observed in large quantities for two
decades in the warm and dense hot cores of massive protostars
\citep{Blake1987}. They have received renewed interest in the last few
years after the detection of abundant COMs in solar-type protostars,
specifically in hot corinos \citep{Cazaux2003,Bottinelli2007}, and in
the clouds of the Galactic Center \citep{RequenaTorres2006}.
Astrochemical models have shown that many COMs cannot be produced
efficiently in the gas phase. \citet{Horn2004}, for example, predicted
methyl formate abundances to be less than $10^{-10}$ by considering
formation pathways only in the gas phase, but this molecule has
been detected with abundances up to $10^{-6}$ (see references in the
caption of Fig. \ref{methylformate_vs_methanol}). On the other hand,
several observational and experimental works have highlighted the
catalytic behaviour of interstellar grains for chemical reactions.
First, infrared (IR) observations of protostellar sources carried out with space
or ground-based telescopes have shown that formaldehyde and methanol
can be significant components of the grain mantles with fractional
abundances up to 30\% with respect to water in some cases
\citep{Gibb2004, Boogert2008}. These mantles are believed to be
formed during the cold and dense prestellar core phase mainly via
hydrogenation and oxydation of CO and O. Second,
several laboratory experiments simulating cold cloud conditions have
confirmed the efficiency of the hydrogenation
processes. \citet{Watanabe2002} for instance have succeeded in
producing solid formaldehyde and methanol at low temperature
($\sim$ $10$ K) via CO hydrogenation reactions in a CO-H$_2$O ice mixture.
Modelling the chemistry on the grain surfaces is therefore crucial if we
aim to understand the formation of COMs on the interstellar
grains. Over the past 30 years, several numerical methods based on
ab initio \citep{Allen1977}, Monte Carlo \citep{Tielens1982,
Charnley1992}, or rate equations \citep{Hasegawa1992, Caselli1998a}
approaches have been developed. All these methods give a macroscopic
description of the problem, meaning that they follow the overall
behaviour of the mantle rather than the single particle. This
method, based on the rate equations, though the least precise from a
physical point of view, is the fastest and allows studies of the
chemical evolution in objects with varying physical conditions as a
function of the time. In this context, time-dependent models have
been developed that predict the abundance of many COMs in hot
cores/corinos by considering two phases: a dense and cold pre-collapse
phase during which the grain mantles with simple hydrogenated species
are formed, and a warm-up phase, caused by the collapse, during which
heavier species, and particularly radicals, can react together on the
grain surfaces before sublimating into the gas phase
\citep{Garrod2006, Garrod2008a, Aikawa2008}. A key point of these
models is therefore the presence of radicals (OH, HCO, CH$_3$O)
trapped in the cold mantles. In these models, the radicals are assumed
to be mainly produced by the UV photodissociation of the neutral
species in the grain mantles. \citet{Garrod2006} tested this
assumption by introducing reactions between OH, HCO, CH$_3$, CH$_3$O
to produce methyl formate, dimethyl ether, and formic
acid. \citet{Garrod2008a} then expanded the chemical network of
reactions between radicals to form other COMs such as ethanol,
glycolaldehyde, or acetic acid. Finally, \citet{Aikawa2008} and
\citet{Awad2010} have included the spatial distribution and evolution
of a collapsing cloud with the aim to estimate the size of typical
hot corinos such as the well-known source IRAS16293-2422.
Although the abundances of the simplest COMs, such as formaldehyde
(H$_2$CO) or methanol (CH$_3$OH), are now quite well predicted,
astrochemical models still fail to reproduce the abundance of more
complex ones, such as methyl formate (HCOOCH$_3$), for example. This
molecule is assumed to be mainly formed on grains via the reaction
between HCO and CH$_3$O during the warm-up phase
\citep{Allen1977,Garrod2006}, a reaction in competition with the
hydrogenation reaction of CH$_3$O, which leads to methanol. Figure
\ref{methylformate_vs_methanol} shows the observed and the predicted
methyl formate to methanol gas phase abundance ratios as function of
the gas phase methanol abundance. The plotted observations
refer to the abundance values estimated in the articles cited in the
figure caption, where the sizes are either directly estimated by
interferometric observations or indirectly by other considerations
for single-dish observations. The plotted model predictions refer
to gas phase abundances. The curves ``Garrod-F" and ``Garrod-S"
have been obtained from Fig. 4 and 6 of \citet{Garrod2008a} respectively,
the curve ``Awad10" comes from Fig. 5 of \citet{Awad2010}, and the
curve ``Laas11" from Fig. 5 of
\citet{Laas2011}. Each point of the curves represents
a different time or temperature (depending on the model). In these
models, the methanol abundance increases with time, except in Awad
et al. (2010) who considered the destruction of the species in the
gas phase. We remark that while the methanol abundances may suffer
of the uncertainty on the source size and H$_2$ column density, the
methanol to methyl formate abundance ratios are almost unaffected
by these uncertainties. Therefore, unless the observed methanol
abundance is always lower than 10$^{-7}$, which is certainly not the
case, no model can reproduce the totality of
observations. Specifically, the observed ratios are roughly
independent of the methanol abundance and are about 0.1 and 1 in hot
cores and hot corinos respectively, regardless of the telescope used
for the observations. State-of-art astrochemical models consequently
underestimate the methyl formate to methanol abundance ratios and
always predict a decreasing ratio with increasing methanol abundance
(because methyl formate is in competition with methanol formation).
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{methylformate_methanol_ratio.ps}
\caption{Gas phase methyl formate to methanol abundance ratio as function of
the gas phase abundance of methanol. Observations of hot corinos with single-dish telescopes
are represented by red crosses: NGC1333-4A by
\citet{Bottinelli2004}, IRAS16293 by \citet{Cazaux2003}, NGC1333-2A
and -4B by \citet{Bottinelli2007}. Observations of hot cores with
single-dish telescopes are represented by blue crosses: G34.3+0.15
by \citet{Mehringer1996}, SgrB2(N) and SgrB2(M) by
\citet{Nummelin2000}, G327.3-0.6 by \citet{Gibb2000}, OMC1 by
\citet{Sutton1995}, G34.3+0.2, SgrB2(N), DR23(OH), W51, Orion
Hot-Core by \citet{Ikeda2001}, AFGL2591, G24.78, G75.78, NGC6334,
NGC7538, W3(H2O), W33A by \citet{Bisschop2007}. Observations of hot
cores with interferometers are represented by blue diamonds:
G34.3+0.15 by \citet{Macdonald1996}, G19.61-0.23 by \citet{Qin2010},
Orion KL, G29.96 by \citet{Beuther2009}, and G47.47+0.05 by
\citet{Remijan2004}. The red curves report the predictions of the
methyl formate to methanol abundance ratio appropriate to the hot
corinos case by Garrod et al. (2008, their Fig. 6) and Awad et
al. (2010, their Fig. 5). The blue curves report the predictions
appropriate to the hot cores case by Garrod et al. (2008, their
Fig. 4) and Laas et al. (2011, their Fig. 5). The arrows represent
the direction of the time in each model and the ticks refer to timescales.}
\label{methylformate_vs_methanol}
\end{figure}
Motivated by this unsatisfactory situation, we developed a new model,
called GRAINOBLE, whose ultimate goal is to simulate the
synthesis of COMs on the grain surfaces. In this article, we focus on
the first phase, namely the formation of the mantle during the dense
and cold phase of the pre-collapse, the prestellar core phase.
Our model differs from most previous published models because it treats
the multilayer and porous structure of the iced mantles. In the
mentioned previous models (Garrod \& Herbst 2006; Garrod et
al. 2008; Aikawa et al. 2008; Awad et al. 2010; Laas et al. 2011),
COMs are synthesized during the warm-up phase thanks to the
increased mobility of radicals, formed on the grain surfaces during
the previous cold phase. The abundance of radicals in the mantles
during the cold phase is, therefore, a key point. One possibility,
explored by the above models, is that the radicals are formed by the
UV photodissociation of frozen species. However, since the impact of
the UV photons on the surface chemistry is poorly understood (see
the detailed discussion in \S 2.1), we explore here the possibility
that radicals are synthesized even without photolysis and are trapped
in the grain mantles, thanks to the multilayer mantle
treatment.
The article is structured as follows. In the next section, we review
the models available in the literature, outlining their limits and pointing
out possible improvement. The concluding paragraph of the section
describes the improvements proposed by our new model GRAINOBLE. In
section \ref{sec:astrochemical-model} we describe the technical
details of GRAINOBLE. Section \ref{sec:results} reports the results of
the new computations, discussing a reference set of parameters and a
large grid of runs where several input parameters are varied. In
section \ref{sec:comp-with-prev} we compare the obtained results with
previous models with the twofold goal to validate our code and discuss
the different results from the different adopted physical
models and conditions. Section \ref{sec:comp-with-observ} compares the
GRAINOBLE predictions with the observations of ices and discusses the
constraints on the various parameters of the model. Finally,
\S\ref{sec:concl-persp} concludes the article, emphasising the main
results and perspectives.
\section{Limits of the existing models and need for improvement}
\subsection{The physics of grain surface chemistry}
\noindent
\textit{a) Multilayer vers bulk chemistry}
One of the main limits of many published models is that they do not
distinguish the chemical processes that occur in the mantle bulk and
on the surface. In these models, species that are buried in the mantle
can keep diffusing in the mantle and reacting with others even if
they are covered by new layers. Furthermore, particles landing on the
surface on top of the grain mantle can react with all particles,
regardless of their depth within the mantle. Thus, heavy
reactive particles such as radicals can continue to react with
landing hydrogen atoms even though the radicals are buried deep in the
mantle bulk. The "bulk chemistry method" may therefore (grossly)
underestimate the abundance of radicals in grain mantles.
In contrast, the difference in the chemical behaviour of the mantle
and of the surface may lead to the formation of non-homogeneous grain
ices as suggested by the IR observations. \citet{Tielens1991}, for
instance, by observing two features in the CO bands,
suggested a polar mixture enriched by water in the bottom of the
mantle covered by a non-polar one mainly composed of CO located on the
outermost layers. In addition, experiments have shown that UV and
cosmic rays irradiation can alter the chemistry of astrophysically
relevant ices composed of water, methanol, and other CO-bearing
molecules even in dark cold cloud conditions \citep{Gerakines1996,
Bennett2007}. However, unlike highly energetic cosmic rays, which
can cross through the entire grain mantle, UV photons only penetrate
a limited number of monolayers, depending on the absorption cross-section
of molecules that constitute the ice \citep[see][for an illustrative discussion]{Gerakines2001}.
Modelling the photolytic processes in ices requires,
therefore, a distinction between the surface and the bulk of the
mantle.
The desorption of grain mantles during the warm-up phase
depends on their chemical composition. By carrying out temperature
programmed desorption (TPD) experiments, \citet{Collings2004} demonstrated
that CO ices desorb at $\sim 20$ K, whereas methanol and water ices
desorb at temperatures higher than 100 K, for instance. Numerical
values of desorption energies were deduced from these experiments and
were incorporated in macroscopic rate equation models. Consequently,
the model of \citet{Garrod2008a} predicts a total desorption of the CO
reservoir at 25 K, whereas more refractory species are desorbed at
temperatures higher than 100 K . Actually, a significant amount of volatile
ices can be trapped with highly polar molecules, such as water, in the mantle
bulk. Thus, the desorption of such species is limited by the gradual evaporation of
the mantle main matrix, which depends on other molecules.
The continuous time random walk (CTRW) Monte-Carlo method, introduced by
\citet{Chang2005} and \citet{Cuppen2005}, strongly differs from the
macroscopic methods described above. This ``microscopic'' model
considers the actual hopping of a discrete number of particles from
one site to another but, unlike previous Monte Carlo methods, it takes
into account the actual position of particles so that the spatial
distribution is also included in the simulation. \citet{Cuppen2007}
and \citet{Cuppen2009} have explained the growth of ice monolayers in
dense clouds and showed a differentiation of the chemical composition
in the mantle. More particularly, \citet{Cuppen2009} showed that
under dense cloud conditions, CO can be trapped in the mantle by the
accretion of new species above it before reacting with other
particles. They also highlighted the survivability of very reactive
particles in the mantle bulk by this process. Unfortunately, this
method can only study simple chemical systems because of the CPU time
required by the Monte-Carlo algorithm, and a direct link with the gas
phase is not possible at present.
\citet{Hasegawa1993b} have attempted to treat separately the chemical
processes within the bulk and on the surface with a three-phase
model. They considered the formation of an inert mantle caused by the
accretion of particles onto it. They found that the differences
between the three-phase and the two-phase models are small for stable
species but become non-negligible for reactive species at long
($\geq 10^5$ yr) timescales. Furthermore, the values of desorption and
diffusion energies used in 1993 are lower than the values measured by
the experiments carried out during the past decade. Because higher
desorption and diffusion energies values tend to decrease the rate of
the chemical processes on grain surfaces, the impact of the
three-phase model on the frozen radical abundances would be
significantly higher with the new values. However, at the time of the submission
of this article, this three-phase model has not been pursued in other works.
\\
\noindent
\textit{b) Porous grain surfaces}
Astrochemical models used for predicting the COM formation have so far
only considered perfectly spherical and smooth grains. However,
observations \citep{Mathis1996}, theoretical studies
\citep{Ossenkopf1993, Ormel2009}, and analyses of solar system bodies
have provided evidence for the fluffy and porous structure of
interstellar grains. For example, \citet{Mathis1996} needed to use
interstellar grains with a substantial fraction of vacuum to fit the
observed extinction curves. Numerical simulations of dust
coagulations carried out by \citet{Ossenkopf1993} from a MRN-like
grain size distribution have shown an increase of the size and the
porosity of the dust in molecular cloud conditions. This fluffy
structure of interstellar grains may alter the surface chemistry
because grains can have micro-pores of small apertures compared to
their volume in which species can be trapped.
Thus, porous grains can increase the reactivity of chemical reactions
by trapping volatile particles such as atomic hydrogen in their pores.
\\
\noindent
\textit{c) Photolysis}
As described in the introduction, the new generation of COMs
formation models predicts that COMs are synthesized during the
warm-up phase thanks to the increased mobility of radicals, formed
on the grain surfaces during the previous cold phase \citep{Garrod2006,
Garrod2008a,Aikawa2008,Awad2010,Laas2011}. The abundance of radicals in the
mantles during the cold phase is, therefore, a key point. In these
works, UV photons from cosmic rays play a major role in synthesising
radicals during this phase. However, as for any model, a series of
inevitable assumptions are adopted to treat the process.
The first important assumption regards the photodissociation rates
on the ices. They are assumed to be the same as those in the gas
phase, computed by \citet{Sternberg1987, Gredel1989}.
However, there are various reasons to think that this
may be a serious overestimate. First, grain surfaces could absorb
part of the UV photon energy, ending up with a lower
dissociation rate. Second, once the molecule is broken, if it is
broken, being not in the gas phase but on a surface with almost no
mobility, the photoproducts may recombine almost instantaneously. In addition, some
products can even have enough energy to sublimate \citep[see for example][]{Andersson2006}.
Thus, the products of the
photodissociation are likely different from those in the gas phase,
and the overall formation rate of the assumed products (and
radicals) may be severely lower than assumed.
A second point is related to the flux of CR induced UV photons in dense clouds.
The exact value is fairly uncertain,
because it largely depends on the primary cosmic ray energy spectrum, and on the
grain extinction cross section, which are also uncertain \citep[see for example][]{Padovani2009}. For example,
\citet{Shen2004} have shown that the typical uncertainty on the low-energy
cosmic ray spectrum leads to a variation of the UV flux of more than one
order of magnitude.
Another important assumption regards the penetration of the UV
photons through the bulk of the mantle. Unfortunately, there exists
almost no observational or theoretical ground for that, and it is
not obvious how realistic this assumption is. Indeed, the
experimental work by \citet{Gerakines2000} suggests that the UV
photons can only penetrate a limited number of layers, depending on
the optical properties of the ice.
Finally, even the branching ratios of species caused by the
photodissociation in the gas phase are very poorly known \citep[e.g,][]{Laas2011}.
In this context, various authors omit the photolysis
in their model \citep[e.g,][]{Chang2007,Cuppen2009}.
\subsection{The computational methods}\label{sec:comp-meth}
The rate equations method allows us to study the evolution of the
chemical composition of grain mantles via one temporal differential
equation for each species. This approach allows us the use of complex
chemical networks that include photolytic processes involving hundreds
of species, and allows us to directly link the gas phase and the grain
surfaces. However, rate equations are based on the evolution of
densities of species in gas-phase and on grain surfaces, which are
average values. Therefore this approach can be inaccurate when the
number of particles on the grain becomes low because of the finite surface
of grains. \citet{Caselli1998a} took into account the discrete nature
of the individual composition of grain mantles by modifiying the
reaction rate coefficients, while \citet{Garrod2008a} went further by
modifying the functional form of the reaction rates.
\citet{Biham2001} and \citet{Green2001} introduced a stochastic method
based on the resolution of the master equations (ME), in which the system
solves the time derivatives of the probability of each discrete mantle
composition. In spite of its good accuracy for grain surfaces, this
method can only be used for small chemical networks since the number
of equations grows exponentially with the number of
species. \citet{Lipshtat2003} have thus introduced another method
based on the time derivatives of the moments to study the formation of
molecular hydrogen, while \citet{Barzel2007} have extended this method
for more complex chemical networks. The comparisons between the
methods introduced here show that the rate equations method tends to
overestimate the reaction rates and the abundance of the reactants,
especially when they are in low quantities.
\subsection{Need for improvement: the GRAINOBLE model}
Although the past few years have seen a huge improvement in the
modelling of grain surface chemistry, models are not yet able to
accurately reproduce observations
(e.g. Fig. \ref{methylformate_vs_methanol}). Therefore, something is
still missing in those models. In the following, we review what are,
in our opinion, the areas where improvements are needed and
possible. Our new grain surface model, GRAINOBLE, has been developed
to include the improvements described below.
\\
\noindent
{\it a) Multilayer structure of the ices}
The first obvious limitation of the Herbst, Garrod, and collaborators class of
models, the present state of the art, is that they do not take into
account the multilayering formation of the ices, which leads to the
differentiation observed in the interstellar ices
\citep{Tielens1991,Pontoppidan2003} that is also predicted by the microscopic
Monte-Carlo (MC) models of Cuppen and collaborators, and to the potential
survival of reactive particles in the mantle bulk. On the other
hand, the microscopic MC treatment is too cumbersome in computer time
and difficult, if not impossible, to apply to model realistic
cases. GRAINOBLE uses the rate equation method developed by Herbst and
collaborators (which is at the base of the Garrod models) in a
way that permits us to follow the multilayering structure of the
ice. Therefore, it benefits from the advantages of the rate
equation method
(computer speed) and of the microscopic MC (multilayer) approach. \\
\noindent
{\it b) Porosity of the grains}
Laboratory experiments, numerical simulations of grain coagulation,
mantle formation, and astronomical observations all demonstrate that
interstellar grains are porous \citep{Jones2011}. As previously
discussed, the presence of pores creates traps for the molecules and
atoms on the grain surfaces, leading to an enhancement of reactivity
on the grains. GRAINOBLE includes pores, following the treatment of
\cite{Perets2006}. The number and area of the pores are treated
as parameters (section \ref{sec:astrochemical-model}).\\
\noindent
{\it c) Multiparameter approach}
Many parameters in the microphysics of the problem have highly
uncertain values, like the various activation barriers of key chemical
reactions, or the barriers against diffusion of adsorbed
particules. Additionally, several macroscopic parameters are also
either uncertain or vary depending on the astronomical object. In
particular, it is worth mentioning that the densities and the
temperatures of the prestellar cores depend on their age and their
surroundings.
\section{The GRAINOBLE model}\label{sec:astrochemical-model}
\subsection{General description}
Following the work of \citet{Hasegawa1992}, four main processes
occuring on grains are taken into account.
\begin{itemize}
\item [1)] Gas-phase particles can accrete onto the grains,
which are considered spherical. For a given gas-phase species $i$, the
accretion rate is a function of the thermal velocity of the
gas-phase species $v(i)$, of the cross section of the grains $\sigma(a_d)$, of the density of
the grains $n_d(a_d)$ where $a_d$ is the grain diameter, and of the
sticking coefficient $S(i)$. We assumed the sticking coefficient
estimated by \citet{Tielens2005} for atomic hydrogen and a sticking
coefficient equal to 1 for heavier particles, based on the TPD
experimental work of \citet{Bisschop2006}. For dark cloud
conditions, particles constituting the ice are bound mainly via the
physisorbed van de Waals interactions.
\item [2)] Once the particles are stuck onto the grains, they can
diffuse along the surface via thermal hopping according to the
Boltzmann law, which gives the hopping rate $R_{hop}$. Following
experimental work of \citet{Katz1999}, we neglected tunelling diffusion.
\item [3)] Physisorbed particles can react only via the
Langmuir-Hinshelwood \citep{Hinshelwood1940} process, in which two
species can react when they meet in the same site. The reaction rate
$R_{r}$ is given by the product between the diffusion rate (i.e, the number
of times per second that a species sweeps over a number of sites equal to
the number of sites of the layer) and the
probability of the chemical reaction, which is a function of its
activation energy.
\item [4)] Surface species can desorb only via thermal processes,
\textit{i}) the ``classic" thermal process caused by the thermal balance
of the grain, \textit{ii}) the cosmic ray induced desorption
process in which cosmic rays heat the grains, as suggested by
\citet{Leger1985}, whose heating rate was computed by
\citet{Hasegawa1993a}. The sum of these two thermal desorption rates gives
the total evaporation rate $R_{ev}$.
\end{itemize}
A full list of the symbols used in this work is provided in the appendix.
\subsection{Porosity}\label{sec:porosity}
To model the impact of the porous structures of interstellar grains,
we introduce two types of sites: the \textit{non-porous}, and the \textit{porous}
sites, following \cite{Perets2006}.
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{schema_pores.eps}
\caption{Schematic view of a portion of the mantle layer as modelled by
GRAINOBLE. All pores are assumed to be square, to have the same
size, and no walls (see text). The edges sites, that connect the
non-porous surface and the pores are depicted in blue. In this example, each
pore is constituted by 16 sites, of which the 12 blue sites are edge sites, and the
fraction of area occupied by the pores is 0.1.}
\label{schema_pores}
\end{figure}
The {\it non-porous} sites form a smooth surface that is perfectly regular,
where particles interact directly with the gas phase. The gas phase
species can accrete onto the grain, and the adsorbates can desorb into
the gas phase.
The {\it porous sites} correspond to sites where no direct
accretion and desorption are possible. They are filled up only by
diffusion from non-porous sites at the edge of the pores.
The porous sites cover a fraction $F_{por}$ of the grain surface. For
the sake of simplicity, all the pores are assumed to be square and to
have the same size, constituted by $N_{pore}$ sites, where $N_{pore}
\geq 4$. The fraction of the grain surface occupied by the pores,
$F_{por}$, is poorly known because, to date, no experimental
measurements or numerical estimates have been attempted. Therefore
we ran models with four values for $F_{por}$: 0, 0.3, 0.6, and 0.9,
to cover a large range of possibilities,
The fraction of the edge porous sites, $F_{ed}$, on the grain depends
then on $F_{por}$ and $N_{pore}$ as follows:
\begin{equation}
F_{ed} = 4 \cdot F_{por} \frac{\sqrt{N_{pore}} - 1}{N_{pore}} .
\end{equation}
A posteriori, the actual value of $N_{pore}$ does not influence the results
significantly because what matters is mostly the fraction
of area occupied by the pores and not the size or the number of
pores. In the following, we assumed $N_{pore}=9$.
The approach that we used for modelling the grain porosity is sketched
in Fig. \ref{schema_pores}.
\subsection{Multilayer approach}\label{sec:multilayer}
In order to distinguish the chemistry on the surface and in the bulk
of the mantle, we used a multilayer approach. In practice, the code
follows the chemical processes on a single layer, the one at the
surface of the mantle. In this layer, particles can accrete, diffuse,
react with each other and desorb, depending on whether they are
on non-porous or porous sites. The mantle layer growth is treated at
the same time as the chemistry. Specifically, at each time $t$, the
chemistry is followed by solving the set of differential equations
described in \S 3.9 between $t$ and $t+ \Delta t$ and a new number
of particles on the grain is computed at $t+ \Delta t$. The layer is
considered chemically inert as soon as the number of particles that
are on the layer is equal to the number $N_s(a_d)$ of (porous plus
not-porous) sites of the layer, where $N_s(a_d)$ is given by the
ratio between the grain surface area and the site area $d_s^2$ (see \S
\ref{sec:part-surf-surf}). Because the porous sites are populated
only via diffusion (\S 3.2), they fill up more slowly than the
non-porous sites. It is therefore possible that the layer is considered
inert when some pores are still empty. In this case, the number of
non-porous sites is “artificially” increased. Although this is
somewhat arbitrary, a posteriori this assumption has almost no
impact on the results, as described in \S 4.3. When the layer
becomes inert, the code memorizes its composition and a new and
reactive layer is started. Note that no chemical exchange is
allowed between layers. Finally, the code also takes into account
the growth of the grain size and of the number of sites of the new
layer by assuming that the thickness of the layer is equal to the site
size $d_s$. All layers have, therefore, the same thickness, equal
to the site size. Figure \ref{schema_layer} gives a schematic view
of this approach.
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{schema_layers.eps}
\caption{Schematic view of the multilayer approach adopted by the
GRAINOBLE code. The inner layers of the mantle are chemically inert,
whereas the layer at the surface is chemically reactive (see text).}
\label{schema_layer}
\end{figure}
\subsection{Chemical network}\label{sec:chemical-network}
We adopted the network based on the work by \citet{Allen1977} and
\citet{Tielens1982}, updated for the new theoretical and experimental
results and adapted to model the formation of the main constituents of
intestellar ices. As previously mentioned in the introduction,
observations of ices have shown that grain mantles are mainly composed
of O- and CO- bearing molecules such as water, CO, CO$_2$, or
methanol. Molecules such as NH$_3$ and CH$_4$ produced by the
hydrogenation of N and C can also be found, but only in smaller
quantities \citep{Gibb2004}. In this work, we therefore neglected the
formation of ammonia and methane. In addition, for the
reasons explained in \S 2.1, we also do not consider photolytic
processes. Our model, therefore, considers the formation of stable
species (e.g, formaldehyde and methanol) and radicals (e.g,
HCO, and CH$_3$O) via hydrogenation of frozen CO. Note that we did
not consider CH$_2$OH because, in absence of photolytic processes, it is
formed via the hydrogenation of H$_2$CO. Theoretical studies have
shown that the reaction H$_2$CO + H $\rightarrow$ CH$_2$OH has an
activation barrier (10.5 kcal/mol) more than twice the activation
barrier leading to CH$_3$O (4--6 kcal/mol) \citep{Sosa1986, Walch1993, Woon2002}.
We therefore neglected CH$_2$OH in the present study.
For the H$_2$O formation, we adopted the new reaction scheme based on
the experimental works by \citet{Dulieu2010}, \citet{Ioppolo2010} and
\citet{Cuppen2010}, who estimated the energy barriers of several
reactions.
The recombination reaction of hydrogen atoms, forming molecular hydrogen, is also
taken into account. Since H$_2$ has a low binding energy with respect to water ice
\citep{Tielens1987, Perets2005}, and since its reaction of formation releases
chemical energy \citep{Hollenbach1970}, we assumed
that the H$_2$ molecules are desorbed in the gas phase as soon as they are formed.
Finally, IR observations have shown that
CO$_2$ may be a major component of ices with abundances reaching 30 \%
with respect to to water in some cases \citep{Dartois1999, Pontoppidan2008}. However, the
formation pathways of carbon dioxide, either by a direct surface route
or by UV photolysis, in such cold conditions are still not well
understood. Accordingly, we decided to neglect the formation of this
molecule.
Table \ref{table_energies} lists the chemical species considered in
our network with their binding energies relative to an
amorphous water ice substrate. Indeed, interstellar grains located in
prestellar-cores are already covered by a polar ice mainly composed of
water, formed during an earlier phase when atomic abundances are still
high, before the depletion of CO.
The network used for the grain surface chemistry is reported in Table
\ref{chemical_network}.
\begin{table}[htp]
\centering
\caption{List of species and binding energies assuming that the ice
matrix is mainly water ice.}
\begin{tabular}{c c c}
\hline
\hline
Species & $E_b$(water ice) (K) \\
\hline
H & 450 \tablefootmark{a} \\
O & 800 \tablefootmark{b} \\
CO & 1150 \tablefootmark{c} \\
OH & 2820 \tablefootmark{d} \\
H$_2$O & 5640 \tablefootmark{e} \\
O$_2$ & 1000 \tablefootmark{c} \\
O$_3$ & 1800 \tablefootmark{f} \\
HO$_2$ & 5570 \tablefootmark{c} \\
H$_2$O$_2$ & 6300 \tablefootmark{f} \\
HCO & 1600 \tablefootmark{f} \\
H$_2$CO & 2050 \tablefootmark{f} \\
CH$_3$O & 5080 \tablefootmark{f} \\
CH$_3$OH & 5530 \tablefootmark{c} \\
\hline
\end{tabular}
\label{table_energies}
\tablebib{$^{(a)}$ \citet{Hollenbach1970}; $^{(b)}$ \citet{Tielens1982}; $^{(c)}$ \citet{Collings2004}; $^{(d)}$ \citet{Avgul1970}; $^{(e)}$ \citet{Speedy1996}; $^{(f)}$ \citet{Garrod2006}}
\end{table}
\begin{table}[htp]
\centering
\caption{Surface reactions and activation energies.}
\begin{tabular}{c l c}
\hline
\hline
Number & Reaction & $E_a$ (K) \\
\hline
1 & O + O $\longrightarrow$ O$_2$ & 0 \\
2 & O + O$_2$ $\longrightarrow$ O$_3$ & 0 \\
3 & H + H $\longrightarrow$ H$_{2,gas}$ & 0 \\
4 & H + O $\longrightarrow$ OH & 0 \\
5 & H + OH $\longrightarrow$ H$_2$O & 0 \\
6 & H + O$_2$ $\longrightarrow$ HO$_2$ & 0 \\
7 & H + HO$_2$ $\longrightarrow$ H$_2$O$_2$ & 0 (36\%)\tablefootmark{a}\\
8 & H + HO$_2$ $\longrightarrow$ OH + OH & 0 (57\%)\tablefootmark{a} \\
9 & H + H$_2$O$_2$ $\longrightarrow$ H$_2$O + OH & 1400\tablefootmark{b} \\
10 & OH + OH $\longrightarrow$ H$_2$O$_2$ & 0 \\
11 & H + O$_3$ $\longrightarrow$ O$_2$ + OH & 0 \\
12 & H + CO $\longrightarrow$ HCO & $400$\tablefootmark{c} $< E_a < 2500$\tablefootmark{d} \\
13 & H + HCO $\longrightarrow$ H$_2$CO & 0 \\
14 & H + H$_2$CO $\longrightarrow$ CH$_3$O & $400$\tablefootmark{c} $< E_a < 2500$\tablefootmark{d}\\
15 & H + CH$_3$O $\longrightarrow$ CH$_3$OH & 0 \\
\hline
\end{tabular}
\label{chemical_network}
\tablebib{$^{(a)}$ branching ratios measured by \citet{Cuppen2010}; $^{(b)}$ \citet{Klemm1975}; $^{(c)}$ \citet{Fuchs2009}; $^{(d)}$ \citet{Woon2002}}
\end{table}
\subsection{Gas phase initial abundances} \label{sec:gasphase}
The observed and predicted gas-phase abundance of atomic carbon is much lower
than CO or atomic oxygen abundances. Therefore, we
considered only the accretion of gaseous H, O, and CO onto the grain
mantles.
In dark cloud conditions, the abundance of atomic hydrogen in the gas phase results from a
balance between the cosmic rays ionization of H$_2$ on the one hand
and the accretion on dust grains (which react to form H$_2$ or heavier
iced species) on the other hand. At steady state, the density of H is
therefore given by the ratio between these two processes
\citep{Tielens2005}, as follows:
\begin{equation}
n(H) = \frac{2.3 \cdot \zeta_{CR} \cdot n(H_2)}{2\cdot v(H) \cdot \sigma_d \cdot X_d \cdot n(H_2)} \label{eq_nH}
\end{equation}
where $v(H)$ is thermal velocity of atomic hydrogen. The resulting H density is independent of the density and equal
to 1.2 cm$^{-3}$ at 10 K, for a cosmic rays ionization rate $\zeta$ of
$3 \cdot 10^{-17}$ \citep{Caselli1998b}, a grain abundance $X_d$ of $1.33
\cdot 10^{-12}$ relative to H nuclei and for a grain radius $r_d$ of 0.1
$\mu$m ($\sigma_d = \pi \cdot r_d^2$), giving
$n_{d} \cdot \sigma_{d}/n_H = 4.2 \cdot 10^{-22}$ cm$^{2}$. Observations
and astrochemical models show that CO is a very stable species and the
most abundant molecule after H$_2$ in dark clouds. Following the
estimates of \citet{Frerking1982}, we assumed an initial abundance
of CO of $4.75 \cdot 10^{-5}$ with respect to H nuclei. This abundance
decreases with time because of the freeze-out onto the grain mantles.
Finally, the abundance of atomic oxygen in the gas phase remains
poorly constrained. Observations of the [OI] 63 $\mu$m line obtained
by \citet{Caux1999, Vastel2000, Lis2001} with the Infrared Space
Observatory (ISO) have measured abundances higher than $10^{-4}$,
suggesting that all the gaseous oxygen is in atomic form.
But astrochemical models predict an atomic oxygen abundances of
$6 \cdot 10^{-5}$ or less. The initial abundance of O is therefore
considered as a free parameter and can vary between $2 \cdot 10^{-4}$
and $2 \cdot 10^{-5}$.
\subsection{Physical conditions}\label{sec:physical-conditions}
As supported by observational and theoretical arguments, the physical
properties of prestellar cores vary substantially with the age and the
mass of the object \citep[see][]{diFrancesco2007, Bergin2007}. The
densities $n_H$ evolve typically from $\sim 10^4$ to $\sim 10^6$
cm$^{-3}$, while the gas kinetic temperature (equal to the grain
temperature in most cases) varies between 7 and 20 K,
depending on the type of the object. Indeed, massive stars are
suspected to be formed from warmer prestellar cores than low mass
ones. Consequently, we included several values of the density and the temperature, listed
in Table \ref{table_grid}, in our parameter grid.
Additionally, as shown by numerical simulations \citep{Ossenkopf1993,
Ormel2009} and observations \citep{Stepnik2003, Steinacker2010}, the
coagulation of grains occuring in molecular clouds and in prestellar
cores tend to increase their size. That is why the grain size distribution,
which is assumed to follow a broad nearly power-law in diffuse interstellar
clouds \citep{Mathis1977}, tends to evolve with time, the bigger
grains dominating most of the dust mass. We assumed all
interstellar grains to have the same
diameter $a_d$, and we considered three values: 0.1, 0.2, and 0.3 $\mu$m respectively.
The grain abundance $X_d$ changes in accordance with the grain size,
to keep a dust-to-gas mass ratio of 1 \%.
\subsection{Other key parameters of GRAINOBLE}\label{sec:part-surf-surf}
In addition to the above described parameters, astrochemical models of
grain chemistry are based on several other parameters related to the
particle-surface, surface particle-particle interaction and surface
properties. We list below three additional key parameters of the
GRAINOBLE model whose values are poorly constrained
because of the difficulties involved in measuring or computing them.
We considered them as free parameters in a
range defined by the different values found in the literature. \\
\noindent
{\it a)} Diffusion energy to binding energy ratio $E_d/E_b$
Theoretical calculations of barriers against diffusion for physical
adsorption $E_d$ on perfectly smooth surfaces have been carried out by
\citet{Jaycock1986} and have shown that $E_d$ is typically about 30 \%
of the binding energy $E_b$. However, ``real" surfaces, which show
defects, irregularities, or steps, tend to increase this ratio
\citep{Ehrlich1966}. Accordingly, several experimental studies have been
carried out to constrain the diffusion barrier on
astrophysically relevant surfaces. By fitting experimental formation
of molecular hydrogen on refractory bare surfaces and on amorphous
water ices with a rate equation model, \citet{Katz1999} and
\citet{Perets2005} experimentally found high $E_d/E_b$ ratios (0.77
and $\sim 0.85$ respectively). However, \citet{Collings2003},
\citet{Ulbricht2002}, and \citet{Matar2008} estimated lower energy
ratios (of about 0.5) for CO and D on water ice, and for atomic oxygen
on carbon nanotubes. Clearly, the energy $E_d/E_b$ ratio is
highly uncertain and strongly depends on the composition and the
structure of the substrate. Three values of the $E_d/E_b$ energy ratio
have been considered here: 0.5, 0.65, and 0.8.\\
\noindent
{\it b)} Activation energy $E_a$
The activation energy values of the CO and H$_2$CO hydrogenation
reactions are also very uncertain. Theoretical studies of these
reactions on an icy mantles via calculations of quantum chemistry
carried out by \citet{Woon2002} and \citet{Goumans2011}, showed that
the activation energies are about 2000-2500 K, and depend slightly on the
presence of water molecules around the reactants. In contrast to these
theoretical values, experiments show that these reactions take place
at low temperature \citep{Watanabe2002}. \citet{Fuchs2009}, by
modelling the experiments of a hydrogen deposition on a CO ice with a
CTRW method, found values of about 400-500 K for both reactions. We
therefore consider these barriers as a free parameter. Three values of
activation energies, assumed to be equal for the H+CO and H+H$_2$CO
reactions, have been considered: $E_a = 400, 1450, 2500$ K.\\
\noindent
{\it c)} Site size $d_s$
The surface density of sites $s$, i.e. the number of sites per surface
unit given in sites x cm$^{-2}$, depends on the structure of the
surface and its composition. Most astrochemical models assume a
value of $s = 10^{15}$ cm$^{-2}$ which has been measured
experimentally by \citet{Jenniskens1995} for a high-density amorphous
water ice. By assuming that the distance between two sites $d_s$ is
constant along the surface, $d_s$ can be easily deduced from the
surface density $s$ ($d_s = 1/\sqrt{s}$), and is equal to $3.2$
$\AA$. However, \cite{Biham2001} found a value of $s = 5 \cdot
10^{13}$ cm$^{-2}$ ($d_s = 14$ $\AA$) for an amorphous carbon surface
and \citet{Perets2006} assumed a value of $s = 5 \cdot 10^{15}$
cm$^{-3}$ ($d_s = 1.4 $ $\AA$) in their model of porous grains. We
therefore treat $d_s$ as a free parameter, with values equal to 1.4,
4.2 and 7 $\AA$. \\
\subsection{Multi-parameter approach} \label{sec:multiparam}
To summarize, the GRAINOBLE model depends on eight key parameters
presented in the previous paragraphs and listed in
Tab. \ref{table_grid} along with the range and values considered in
this study for each of them. The values reported in bold are used in
the ``reference" model (\S \ref{sec:results}). They correspond to the
average physical conditions of prestellar cores (density, temperature,
and grain size), to the average values found in the litterature (site size,
diffusion energy to binding energy ratio, initial abundance of atomic oxygen),
to the values of parameters used by most of previous astrochemical models
(porosity factor) or to the most recent values
measured by experiments (activation energy).
In total, we ran a grid of 17496 models varying these eight free parameters.
\begin{table}[htp]
\centering
\caption{List of the key free parameters of GRAINOBLE and the value range.}
\begin{tabular}{c c}
\hline
\hline
Parameter & Values \\
\hline
Density $n_H$ & $10^{4}$ - $\mathbf{10^5}$ - $10^{6}$ cm$^{-3}$ \\
Temperature $T_g = T_d$ & 10 - \textbf{15} - 20 K \\
Initial oxygen abundance $X$(O)$_{ini}$ & $2 \cdot 10^{-5}$ - $\mathbf{6 \cdot 10^{-5}}$ - $2 \cdot 10^{-4}$ \\
Grain size $a_{d}$ & 0.1 - \textbf{0.2} - 0.3 $\mu$m \\
Energy ratio $E_d/E_b$ & 0.5 - \textbf{0.65} - 0.8 \\
Porosity factor $F_{por}$ & \textbf{0} - 0.3 - 0.6 - 0.9 \\
Site size $d_s$ & 1.4 - \textbf{4.2} - 7 $\AA$ \\
Activation energy $E_a$ & \textbf{400} - 1450 - 2500 K \\
\hline
\end{tabular}
\tablefoot{Bold values mark the values adopted in the reference
model (see text, \S \ref{sec:results}).}
\label{table_grid}
\end{table}
\subsection{Computational aspects} \label{sec:equadiff}
The code solves three sets of differential equations: the first one
describes the evolution of the densities of species (other than H, whose density is
given by eq. \ref{eq_nH}) in the gas phase,
the other two describe the chemical composition of the mantle
outermost layer on the non-porous surface and within the pores. The
equations giving the evolution of the chemical composition of the
mantle layer do not depend on the grain size and are expressed in
monolayers/sec (MLs/s) where a monolayer is given by the number of particles
of the considered species divided by the number of sites of a layer.
The equations are the following:
\begin{eqnarray}
\frac{dn_{g}(i)}{dt} = &-& S(i) \cdot v(i) \cdot n_{g}(i) \cdot \sigma(a_{d}) \cdot n_{d}(a_{d}) \nonumber \\
& + & R_{ev}(i) \cdot P_{np}(i) \cdot N_{s}(a_d) \cdot n_d(a_d)
\label{ed_gas}
\end{eqnarray}
\begin{eqnarray}
\frac{dP_{np}(i)}{dt} = & & \frac{1}{4} \cdot S(i) \cdot d_{s}^2 \cdot v(i) \cdot n_g(i) - R_{ev}(i) \cdot P_{np}(i) \nonumber \\
& - & \frac{F_{ed}}{F_{np}} \cdot R_{hop}(i) \cdot P_{np}(i) + \frac{F_{ed}}{F_{por}} \cdot R_{hop}(i) \cdot P_{por}(i) \nonumber \\
& + & \sum_{i_f} R_{r,ext}(i_f) \cdot P_{np}(i_{r1}) \cdot P_{np}(i_{r2}) \nonumber \\
& - & \sum_{i_d} R_{r,ext}(i_d) \cdot P_{np}(i_{r1}) \cdot P_{np}(i_{r2})
\label{ed_ext}
\end{eqnarray}
\begin{eqnarray}
\frac{dP_{por}(i)}{dt} = &+& \frac{F_{ed}}{F_{np}} \cdot R_{hop}(i) \cdot P_{np}(i) - \frac{F_{ed}}{F_{por}} \cdot R_{hop}(i) \cdot P_{por}(i) \nonumber \\
& + & \sum_{i_f} R_{r,in}(i_f) \cdot P_{por}(i_{r1}) \cdot P_{por}(i_{r2}) \nonumber \\
& - & \sum_{i_d} R_{r,in}(i_d) \cdot P_{por}(i_{r1}) \cdot P_{por}(i_{r2})
\label{ed_in}
\end{eqnarray}
where $P_{np}(i)$ and $P_{por}(i)$ are the surface populations of the
species $i$ (in MLs) on the non-porous surface and within the pores
respectively, and $n_{g}(i)$ its gas-phase density. The first and
second terms in equation \ref{ed_gas} describe the accretion rate and
the evaporation rate. The first and second
terms of equation \ref{ed_ext} describe the accretion and the
evaporation rates. The third and fourth terms describe the exchange
rate between the non-porous surface and the pores via diffusion.
The fifth and sixth terms describe the production and
the destruction rates of $i$ via reactions on the non-porous
surface.
Similarly, the first and second terms of equation \ref{ed_in} describe
the exchange rate between the non-porous surface and the pores via diffusion.
The third and fourth terms describe the production and
the destruction rates of $i$ via reactions in the pores.
Note that the evolution of composition within all the pores can be treated with a
single equation because all pores are assumed to have the same
size. The equations giving the evolution of each pore can be deduced
from equation \ref{ed_in} by multiplying it with a simple factor of
proportionality.
\subsection{What the model does not consider}
We aim to study the impact of the multilayer behaviour and
the porous structure of the grains on the mantle formation and
its chemical composition. To this end, we considered thermal processes
only. Therefore we did not consider the
desorption processes caused by exothermic surface reactions
\citep{Garrod2007}, and the photolytic desorption
\citep{Oberg2009a, Oberg2009b, Arasa2010} in spite of their relative
possible importance. Neither did we account for the
photodissociation processes on the ice in
this version of GRAINOBLE owing to the reasons given in Sect. 2.1.
\section{Results}\label{sec:results}
We ran a grid of 17496 models, varying the eight
free parameters listed in Table \ref{table_grid}. In this section, we
describe the results of our model. We start by describing the results
relative to the reference model, and then we describe how the results
change when each of the parameters changes. The analysis is based on
the distribution of the obtained mantle abundances when all
parameters are varied, except for the parameter under consideration. This
allows us to understand if and by how much the considered parameter
influences the results. The plot of the multilayer and bulk distributions
is shown in section \ref{sec:multilayer-vs-bulk},
while the distributions of other parameters (using the multilayer approach only)
are shown in the appendix.
Finally, Table \ref{summary_grid} summarises the main effects on the
resulting mantle abundances caused by each parameter.
\subsection{The reference model}\label{sec:reference-model}
This section describes the results obtained using the reference set of
parameters (the boldface values in Table \ref{table_grid} ). Below
we refer to it as the ``reference model''. Figure
\ref{time_Xgrain} presents the evolution with time of the abundance of
each species in the gas phase and in the grain mantles.
\begin{figure*}[htp]
\centering
\includegraphics[width=180mm]{new_X_all.ps}
\caption{Abundance of species in interstellar grain mantles (solid)
and in gas phase (in dashed) as a function of time for the
reference model (boldface values in Table
\ref{table_grid}). \textit{Left panel:} the most abundant stable
species. \textit{Middle panel:} reactive species. \textit{Right
panel:} less abundant stable species.}
\label{time_Xgrain}
\end{figure*}
The evolution of the chemical composition of the mantle depends on the
initial gas phase abundances. In the reference model, O and CO
abundances are approximately five times higher than the H abundance.
Consequently, H atoms cannot hydrogenate all O and CO molecules that
accrete onto the mantles. Besides, because the CO hydrogenation
reaction has a barrier of 400 K, H atoms react preferentially with O
and its products via barrierless reactions, ending up in water
molecules. In $2 \cdot 10^5$ years, 80 \% of the CO reservoir is
trapped within the mantle bulk, whereas $\sim 60$ \% of O atoms are
contained in water. The remaining O atoms are shared between O$_3$
(ozone), OH and H$_2$O$_2$ (hydrogen peroxyde). Indeed, the high O to
H initial gas phase abundance ratio ($\sim 6$) and the relatively high
temperature (15 K) allow a significant amout of O atoms to diffuse on
the surface and to react with each other before meeting hydrogen atoms.
However, gas phase CO and O abundance ratios relative to H gradually
decrease with time (because they freeze-out onto the mantles), increasing
the possibility of hydrogenation reactions. After $\sim 5 \cdot 10^4$
years, the formation rates of formaldehyde and methanol start to
increase while the formation of ozone stops, corresponding to the time
when H, O, and CO have similar abundances. The percentage of ozone,
formaldehyde, and methanol in the ices therefore strongly depends on the
age of the core. Young cores would show high abundances of ozone and
of hydrogen peroxyde with respect to water, whereas older ones would
show high abundances of formaldehyde and methanol.
Another important result of the model is that a significant amount of
radicals is trapped in the bulk because of the multilayer
treatment. Radicals assumed to be the precursors of COMs such as OH,
HCO or CH$_3$O, reach abundances between $\sim 5 \cdot 10^{-6}$ and
$10^{-9}$. OH is more abundant because of its barrierless formation
reaction, and its formation mostly occurs at shorter times.
The evolution of the chemical composition can also be studied
``spatially". Indeed, our multilayer approach allows us to study the
composition of each monolayer within the grain mantle. Figure
\ref{time_ML} shows the evolution of the formation time of each layer
and the evolution of the mantle thickness with time. In the reference
model, a grain mantle of 77 layers is formed in $2-3 \cdot 10^5$
yr. Assuming that the layer thickness is equal to the site size (here
4.2 $\AA$), it gives a total mantle thickness of $\sim 0.04~\mu$m,
namely 32 \% of the grain radius. The first 60 layers are created in
less than $10^5$ yr because they have a fast formation rate (each layer
is formed in less than $\sim 2 \cdot 10^3$ yr). The formation time of
layers increases sharply at $\sim 10^5$ yr, because of the drop of the gas
phase abundances.
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{layers2.ps}
\caption{Formation time of each monolayer (blue line) and mantle
formation time versus its thickness expressed in monolayers (red
line) for the reference model (boldface values in Table
\ref{table_grid}).}
\label{time_ML}
\end{figure}
The differentiation of the composition between the inner and the outer
layers of the mantle can also be seen in Fig. \ref{ML_compo}. Indeed,
``intermediate'' molecules such as O$_3$, which is created from
atomic oxygen, are rapidly formed and are therefore abundant in the inner
layers of the mantle (the first 50 layers), whereas formaldehyde
and methanol are mostly formed later, their fractional composition
showing significant abundances only in the 20 outermost layers. The
abundances of the two main species, CO and H$_2$O, remain relatively
constant throughout the mantle up to the last outermost layers, where
they decrease in favour of H$_2$CO and CH$_3$OH. The composition of
radicals within the mantle follows the same evolution as their
precursor molecules: the HCO abundance stays relatively constant until
the drop in the last layers like CO, while the OH abundance decreases
following the O abundance behaviour. Finally, the CH$_3$O abundance
increases approaching the surface, as H$_2$CO.
\begin{figure*}[htp]
\centering
\includegraphics[width=180mm]{new_poplay_all.ps}
\caption{Fractional composition of each mantle monolayer for the
reference model (boldface values in Table \ref{table_grid}).}
\label{ML_compo}
\end{figure*}
\subsection{Multilayer versus bulk approach}\label{sec:multilayer-vs-bulk}
The introduction of the multilayer approach modifies the chemical
behaviour of grain mantles and thus their overall chemical
composition. Figure \ref{time_bulkvslayer} shows the evolution of the
mantle composition with time for the reference model,
adopting a multilayer (described in sec. \ref{sec:multilayer}) or a bulk (old method described in 2.1) approach, respectively.
The figure shows that at short timescales ($\leq 5\times10^4$ yr),
both approaches give similar abundances for the main species
because the H abundance is initially low compared to the CO and O
abundances (\S \ref{sec:reference-model}). In contrast, O and its
associated ``intermediate'' molecules (OH, O$_2$, and HO$_2$) are more
efficiently burned in O$_3$, H$_2$O$_2$, and H$_2$O in the bulk compared
to the multilayer approach.
At later times ($\geq 5\times10^4$ yr), reactive species, which are
trapped by the multilayer approach, continue to react in the bulk
method. Consequently, in the bulk approach, radicals (OH, HCO, HO$_2$,
CH$_3$O) and ``intermediate'' stable species (O, O$_3$, H$_2$O$_2$,
O$_2$) are totally burned in less than $\sim10^6$ yr. In contrast,
the abundances of these species do not evolve after $\sim3 \cdot 10^5$
yr with the multilayer approach. Finally, the predicted methanol and
formaldehyde abundances substantially diverge in the two approaches
at times longer than $\sim 10^5$ yr, namely the presumed ages of the
prestellar cores. The bulk method predicts a final (at $10^6$ yr)
methanol abundance higher by factor 6 compared to the multilayer
approach, and a formaldehyde abundance lower by a factor higher than 10$^4$.
\begin{figure*}[htp]
\centering
\includegraphics[width=180mm]{new_Xmantle_all.ps}
\caption{Abundance of species in interstellar grain mantles as a
function of time for the reference model (solid lines), the bulk
approach model (\S \ref{sec:multilayer-vs-bulk}; dashed lines), and
the porous grain model (\S \ref{sec:infl-grain-poros}; dotted
lines). \textit{Left panel:} the most abundant stable
species. \textit{Middle panel:} reactive species. \textit{Right
panel:} less abundant stable species.}
\label{time_bulkvslayer}
\end{figure*}
To understand how robust these results are, we considered the whole
grid of models described in \S \ref{sec:multiparam} and built the
distribution of the predicted abundances. Figure \ref{distrib_treat}
shows this distribution for key stable species (CO, H$_2$O, H$_2$CO,
and CH$_3$OH) and the radicals (OH, HCO, and CH$_3$O) for
the bulk and multilayer approaches for two times ($10^5$
and $10^6$ yrs). The plots of Fig. \ref{distrib_treat} tell us that at
short times ($10^5$ yr) the predicted species abundance distributions
are quite similar in the bulk and multilayer approaches, with the
exceptions of H$_2$CO and CH$_3$O, for which the distribution are
slightly different. In contrast, at long times ($10^6$ yr), the
distribution of all radicals are substantially different in the two
approaches, demonstrating that the age is a key parameter. The plots
also provide another important information. For example, considering
the multilayer approach results, CO and OH have quite peaked
distributions and their predicted abundances depend slightly if at all,
on the assumed values of the parameters: they are robust
predictions. On the contrary, H$_2$CO, CH$_3$OH, HCO and CH$_3$O have
broad abundance distributions implying that their predicted abundances
are very sensitive to the values of the other model parameters.
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{distrib_lin_all_treat_new.ps}
\caption{Distribution of the predicted mantle abundance $X_{mantle}$ of the key
species
at $10^5$ (left panels) and $10^6$ yr (right panels).
The thin blue and thick red lines refer to the bulk
and multilayer approach. The distribution has been
built by considering all the $\sim$18000 runs of the grid (\S
\ref{sec:multiparam}).}.
\label{distrib_treat}
\end{figure}
\begin{normalsize}
\begin{table*}[htp]
\caption{Summary of the effect of each free parameter on the surface chemistry and on the distribution of results (shown in the appendix).}
\begin{tabular}{>{\centering}p{2cm}>{\raggedright}p{5cm}>{\raggedright}p{5cm}>{\raggedright}p{5cm}}
\hline
\hline
Parameter & Effect on the surface chemistry & Distributions of abundances for stable species & Distributions of abundances
for radicals\tabularnewline
\hline
Multilayer versus bulk approach & The multilayer approach
- decreases the formation of stable species
- traps reactive species in the mantle & - $X$(CO) $> 10^{-5}$ for
88\% (ML), 50\% (bulk) cases
- $X$(H$_{2}$O) $> 10^{-5}$ for
60\% (ML), 93\% (bulk) cases
- $X$(CH$_{3}$OH) $> 10^{-5}$ for
15\% (ML), 55\% (bulk) cases
& - X(OH) $> 10^{-9}$ for
75\% (ML), 1\% (bulk) cases
- $X$(HCO) $> 10^{-9}$ for
25\% (ML), 2\% (bulk) cases
\tabularnewline
\hline
Porosity & The increase of porosity
- traps volatile and diffusing particles
- increases the formation rate of stable species & - $X$(H$_{2}$CO) $> 10^{-7}$ for
52\% (smooth), 61\% (porous) cases
- $X$(CH$_{3}$OH) $> 10^{-7}$ for
51\% (smooth), 61\% (porous) cases
& - X(OH) $> 10^{-6}$ for
55\% (smooth), 20\% (porous) cases
\tabularnewline
\hline
Density & The increase of density
- decreases the formation of stable species formed by hydrogenation
reactions & - $X$(H$_{2}$O) $>$ 10$^{-5}$ for
2\% (dense) , 85\% (sparse) cases with $X$(H$_{2}$O) $\rightarrow 2\cdot10^{-4}$
- $X$(CH$_{3}$OH) $> 10^{-5}$ for
0\% (dense) , 35\% (sparse) cases with $X$(CH$_{3}$OH) $\rightarrow 9\cdot10^{-5}$
& - $X$(OH) $> \cdot10^{-6}$ for
0\% (sparse), 75\% (dense) cases
\tabularnewline
\hline
Temperature & The increase of temperature
- strongly increases the desorption of H and O
- decreases the rates of hydrogenation reactions
- decreases the abundance of stable species formed by these reactions & - $X$(CH$_{3}$OH) $> 10^{-7}$ for
65\% (cold), 45\% (warm) cases
- $X$(H$_{2}$CO) $> 10^{-7}$ for
60\% (cold), 45\% (warm) cases & - $X$(OH) $> 10^{-6}$ for
55\% (cold), 48\% (warm) cases\tabularnewline
\hline
Grain size
$a_{d}$ & The increase of the grain size
- increases the mantle thickness
- does not influence the integrated mantle composition & Same evolution of distributions for all stable species & Same evolution of distributions for all radicals\tabularnewline
\hline
Initial abundance of atomic oxygen & The increase of $X_{ini}$(O):
- increases the formation of water (and other molecules formed from
reactions involving O)
- slightly decreases the formation of molecules formed from CO & - $X$(H$_{2}$O) $> 2 \cdot 10^{-5}$ for
0\% (low-O), 65\% (high-O) cases with $X$(H$_{2}$O) $\rightarrow 2\cdot10^{-4}$
- No influence is seen for CO, H$_{2}$CO and CH$_{3}$OH & - $X$(OH) $>4\cdot10^{-6}$ for
0\% (low-O), 50\% (high-O) cases with $X$(OH) $\rightarrow 4\cdot10^{-5}$ \tabularnewline
\hline
Diffusion energy & The increase of the diffusion energy
- decreases the diffusion rate of mobile species
- decreases the formation rate of stable species
- slightly increases the survival of radicals & - $X$(H$_{2}$O) $>5\cdot10^{-6}$ for
87\% (fast), 74\% (slow) cases
- $X$(CH$_{3}$OH) $> 10^{-7}$ for
80\% (fast), 30\% (slow) cases
- $X$(H$_{2}$CO) $> 10^{-7}$ for
75\% (fast), 40\% (slow) cases & Same evolution of distributions for all radicals \tabularnewline
\hline
Activation energies & The increase of activation energies
- strongly decreases the formation rate of H$_{2}$CO and CH$_{3}$OH
- strongly decreases the survival of radicals HCO and CH$_{3}$O
- slightly increases the reaction rates of reactions involving O
- slightly increases the formation of water & - $X$(CO) $> 10^{-5}$ for
35\% (low-Ea), 98\% (high-Ea) cases
- $X$(H$_{2}$CO) $>10^{-7}$ for
83\% (low-Ea), 25\% (high-Ea) cases
- $X$(CH$_{3}$OH) $>10^{-6}$ for
80\% (low-Ea), 5\% (high-Ea) cases & - $X$(HCO) $>10^{-10}$ for
90\% (low-Ea), 0\% (high-Ea) cases
- $X$(CH$_{3}$O) $>10^{-10}$ for
70\% (low-Ea), 0\% (high-Ea) cases\tabularnewline
\hline
Site size & The increase of the site sizes
- decreases the formation time of layers
- increases the mantle thickness
- does not influence the integrated mantle composition & Same evolution of distributions for all stable species & Same evolution of distributions for all radicals\tabularnewline
\hline
\end{tabular}
\label{summary_grid}
\end{table*}
\end{normalsize}
\subsection{Porous versus non-porous grains}\label{sec:infl-grain-poros}
The influence of the grain porosity (\S \ref{sec:porosity}) is shown
in Figure \ref{time_bulkvslayer} for the reference set of parameters.
The presence of the porosity in the grain has the effect of slightly
speeding up and increasing (by less than a factor 4) the formation of
formaldehyde, methanol, and CH$_3$O, as well as to increase the
destruction of HO$_2$, O$_2$ and O. The reason is that porous grains
trap H atoms more efficiently, increasing their number on the grains,
but in the end, the H atom numbers is still limited by their
low gas phase abundance (Fig. \ref{time_bulkvslayer}).
Furthermore, the porosity only slightly modifies the distribution of the results
shown in Fig. A.1.a. Therefore, the porosity only enhances
the formation of hydrogenated molecules by a few factors at most,
regardless of the values of other parameters.
\subsection{Influence of physical conditions}
\noindent
{\it Density $n_H$ }
The density plays an important role for two reasons. First, increasing
the density reduces the time needed to form each monolayer (because the
accretion rate is proportional to the square of the density). Second, the ratio
between the number of H atoms and the number of heavy particles (CO
and O) that land on the grain is inversely proportional to the density
(see section \ref{sec:gasphase}). As a consequence, the higher the
density, the lower the abundance of stable species (H$_2$O, H$_2$CO,
CH$_3$OH) created by the hydrogenation reactions, especially at long
times ($\geq10^5$ yr) (Fig. A.1.b). In contrast and for the same
reasons, mantle CO and OH are more abundant for higher density.\\
\noindent
{\it Temperature $T_g = T_d $}
The grain temperature has a moderate influence on the predicted mantle
composition, although the diffusion (and thus the reaction) and the
desorption rates depend exponentially on it. However, the desorption
rate increases much more quickly than the diffusion rate with
temperature (by factors of $10^2 - 10^6$ depending on the $E_d/E_b$
ratio). Consequently, with increasing grain temperature, H atoms have
a higher probability to desorb into the gas phase before encountering
another particle. The final mantle abundance of molecules created by
hydrogenation reactions (H$_2$CO and CH$_3$OH) is therefore lower at any
time at higher temperatures, as shown by Fig. A.1.c.
The mantle abundance
of radicals, on the other hand, is not affected by the grain
temperature.\\
\noindent
{\it Grain size $a_d$}
While the mantle thickness depends roughly quadratically on the grain
size, the final mantle composition depends on it very little. The
reason is that the particle accretion rate (inversely proportional to
the grain size square) and the formation time of each layer compensate
each other. The result is that the mantle composition does not depend
on the assumed grain size, whether 0.1 or 0.3 $\mu$m.
\subsection{Influence of other key parameters of GRAINOBLE}
\noindent
{\it Initial abundance of atomic oxygen $X$(O)$_{ini}$}
Not surprisingly, the initial gas phase abundance of atomic oxygen is an
important parameter for the final mantle water abundance. The higher
the O abundance, the higher the iced H$_2$O.
Less evidently, oxygenation is in competition with hydrogenation. The
initial O abundance also affects the final mantle abundance of
formaldehyde and, even more, methanol, because the elemental C/O ratio
decreases between 0.7 and 0.2 when $X($O$)_{ini}$ increases between
$2\cdot 10^{-5}$ and $2\cdot 10^{-4}$ . At 10$^5$ yr, the predicted
methanol abundance is a factor ten higher when the O abundance is a
factor ten lower.\\
\noindent
{\it Diffusion energy to binding energy ratio $E_d/E_b$}
The diffusion energy is an important parameter in the predicted mantle
abundance of formaldehyde and methanol. Because the diffusion rates depend
exponentially on the diffusion energy, the mantle abundance of stable
species, like formaldehyde and methanol, decreases with increasing
$E_d$. Conversely, the diffusion energy does not affect the mantle
abundance of radicals, because their formation and destruction rates
compensate each other.\\
\noindent
{\it Activation energy $E_a$}
The activation energy of CO and H$_2$CO hydrogenation reactions
strongly influences the formation rates of formaldehyde and methanol
because their reaction rates depend exponentially on $E_a$. Most of the runs
with a low value of $E_a$ (400 K) predict mantle abundances of
formaldehyde and methanol higher than $10^{-6}$ at $\geq 10^5$ yr.
Conversely, higher values of $E_a$ give uniform distributions of
H$_2$CO and CH$_3$OH abundances, between $10^{-18}$ and
$10^{-6}$. Only $\sim 5$\% of the runs using a activation energy
equal to 2500 K predict formaldehyde and methanol abundances higher
than $10^{-6}$. The final abundances of formaldehyde and methanol,
consequently strongly depend on the other model parameters for high $E_a$
values. A significant amount of radicals, like HCO and CH$_3$O, can
survive in the mantle only if the activation energies are low. Runs
with activation energies of 400 K predict radical mantle abundances
between $10^{-10}$ and $10^{-8}$ for both radicals, while runs with
$E_a = 2500$ K predict abundances less than $\sim 10^{-11}$. Not
surprisingly, the activation energy of CO and H$_2$CO hydrogenation
reactions are, therefore, critical parameters for the survival of
radicals in the mantle. \\
\noindent
{\it Site size $d_s$}
The reaction rates and the formation time of each monolayer are both
functions of the site size $d_s$: the former is proportional to
$d_s^{2}$, whereas the latter is proportional to $d_s^{-2}$.
Therefore, the two processes, which are in competition, cancel each
other, and consequently, the total mantle abundance of stable species
and radicals do not sensitively depend on the size of sites.
However, the thickness of the grain mantle strongly depends on the site
size. Indeed, the maximum number of monolayers is given by the
ratio between the number of particles on the grain $N_{part}$ and the
number of grain sites $N_s$ (proportional to $d_s^{-2}$). Furthermore,
if we assume that the thickness of a layer is equal to the
distance between two sites, the thickness of the mantle is proportional
to $d_s^3$. Thus, the bigger are the sites, the faster the grains grow.
Consequently, the depletion is more efficiently on grains with big sites,
increasing the abundance of methanol and decreasing the abundance of CO
for example.
\subsection{Concluding remarks}
Based on the previous paragraphs, we can identify three classes of
parameters, depending on their influence on the results:
\begin{itemize}
\item 1) The parameters that have a {\it strong} influence on the
mantle composition are the approach to model the chemical behaviour
(multilayer versus bulk), the density, the diffusion energy and
the activation energy of CO and H$_2$CO hydrogenation
reactions. Modifying the values of these parameters within their
chosen range drastically changes the distribution of the predicted
mantle abundances, both in the shape and in the values.
\item 2) The grain porosity and the temperature as well as the initial
abundance of atomic oxygen have only a {\it moderate} impact on the
predicted mantle composition.
\item 3) The grain and the site sizes have a {\it negligible}
influence on the total mantle composition. However, these parameters
strongly modify the ice thickness, which could have an impact on
photolytic processes and for the desorption time of the mantle
during the warm-up phase in protostar envelopes.
\end{itemize}
\section{Comparisons with previous microscopic models}\label{sec:comp-with-prev}
The Monte-Carlo continuous time random walk (CTRW) method was
introduced by \citet{Chang2005} and \citet{Cuppen2005} for the
hydrogen recombination system, and was then extended to more complex
networks in \cite{Chang2007}, \citet{Cuppen2007},
\citet{Cuppen2009}. The natural output of this model is the multilayer
structure of the mantle. This section compares our results with the
model of \citet{Cuppen2009} with the twofold aim of validating the
GRAINOBLE code and of highlighting the differences between the two
models.
\subsection{Validation of the GRAINOBLE code}
\citet{Cuppen2009} considered the accretion of H, CO, and H$_2$ on
smooth surfaces as described by \citet{Cuppen2005}, leading to the
formation of formaldehyde, methanol, and their associated radicals
HCO and CH$_3$O via the Langmuir-Hinshelwood and the Eley-Rideal mechanisms.
A major difference between GRAINOBLE and the Cuppen et al. model is
that they consider an $\alpha$-CO ice, whereas we assumed that the
bulk of the mantle is formed by iced water. To reproduce the
interactions of molecules within an $\alpha$-CO ice, we modified the
set of binding energies in our model accordingly. Second, the CTRW
model takes into account the individual interaction energies between
each particle, while our model only considers the total binding energy
between the adsorbate and the substrate. To simulate their approach,
we assumed for each species $i$ a binding energy equal to fourteen
times the interaction between the species $i$ and a CO molecule
($E_b(i) = 14 E_{i-CO}$), corresponding to the interaction between an
adsorbate and a porous multi-layered CO ice. The binding energy of H
is equal to 450 K, while the binding energy of CO is 880 K. The
diffusion to desorption energy ratio is assumed to be 0.8, as in
\citet{Cuppen2009} for a layer that is initially smooth. The
activation energies for CO and H$_2$CO hydrogenation reactions are
those measured by \citet{Fuchs2009}, and slightly depend on the grain
temperature.
To validate our code with respect to the model of \citet{Cuppen2009}, we run our model
with a density of $10^5$ cm$^{-3}$ and considered the results at
$2\times10^5$ yr, for the following four cases (as in Part 3.3 of
\citet{Cuppen2009}): a) grain temperature 12 K, $n(H)/n(CO)$=1 and
$n_d = 2\times 10^{12} n_h$ ; b) as a) but $n_d = 1\times 10^{12}
n_h$; c) as a) but temperature 15 K; d) as b) but $n(H)/n(CO)$=0.5.
Figure \ref{results_cuppen} shows the chemical composition of each
monolayer for the four cases, for porous grains ($F_{por} = 0.7$: this
is the value that agrees best for our and Cuppen's
model). The comparisons between Figure 4 of \cite{Cuppen2009} and our
Figure \ref{results_cuppen} shows that the two codes give very similar
results, which validates our code.
\begin{figure}[htp]
\centering
\includegraphics[width=88mm]{new_multiplot_poplay_stable.ps}
\caption{Chemical composition of the mantle layers as a function of
the monolayers at a time of $2 \cdot 10^5$ years, for a
density $n_H = 10^5$ cm$^{-3}$, as in \citet{Cuppen2009}: see
text): CO (red), formaldehyde (blue) and methanol (green). The four
panels refer to the four cases of Fig.4 of \citet{Cuppen2009} (see
text): a) grain temperature 12 K, $n(H)/n(CO)$=1 and $n_d = 2\times
10^{12} n_h$ ; b) as a) but $n_d = 1\times 10^{12} n_h$; c) as a)
but temperature 15 K; d) as b) but $n(H)/n(CO)$=0.5.}
\label{results_cuppen}
\end{figure}
\subsection{Differences between the two approaches}
We highlight a few critical differences between the model of
Cuppen and ours.
Unlike our multilayer approach, the CTRW method allows us to model the
segregation effects. Heavier particles can agglomerate together to
form islands where they can be fixed onto the grains more
efficiently. Our multilayer method is a macroscpic approach where
segration effects are not taken into account. However, the pores can
play the same role as the islands, because particles that enter into
the pores cannot desorb either. This is supported by the fact that our
porous grains case agrees best with Cuppen's model.
An important difference between ours and Cuppen et al.'s model is the
inclusion of the water ice formation. Because Cuppen et al. did not
consider the formation of water, their ices are formed by pure
$\alpha$-CO ice. Consequently, the interaction energies are lower than
those used in our model. Therefore in Cuppen's
model ices can form only at temperatures lower than 16-18 K, whereas in
our model iced CO can still survive until 21-23 K. In our model,
then, the formation of formaldehyde and methanol can occur in a wider
range of temperatures than in \citet{Cuppen2009}. Consequently, the
abundance of formaldehyde and methanol formed on grain mantles decreases
much more efficiently with the increase of the temperature in Cuppen et
al.'s model than in ours. For example, using the same physical parameters, the
CTRW model predicts a decrease of the mantle thickness of 85 \% ($\sim$ 40 to 6
monolayers), whereas our model does not predict any thickness decreases
between 15 and 16.5 K.
\section{Comparisons with observations}\label{sec:comp-with-observ}
\subsection{The observations}
In order to compare our model predictions with observations, we used
the data pertaining to the ices only, because the abundance of the sublimated
mantle species does not necessarly reflect the chemical composition of
their precursor ices. First, it can be altered by reactions in the gas
phase. Second, the different sublimation temperatures may introduce
errors when comparing the abundance ratios of different species, because
they may refer to different regions.
Unfortunately, solid CO, H$_2$CO, and CH$_3$OH have been observed
simultaneously only towards a very small sample of high- and
intermediate-mass protostars \citep{Gibb2004,Pontoppidan2004}. To
increase the statistics and to include low-mass protostars,
too, we therefore restricted the comparison of our model to the
observed solid CO and CH$_3$OH abundances only. Fortunately, {\it a
posteriori}, when considering the model predictions, we found that
using the H$_2$CO would not provide (substantially) more constraints.
Table \ref{obs_CO_CH3OH} reports the compilation of observations that
we used for the comparison.
\begin{table}[htp]
\centering
\caption{Mantle abundances of CO and CH$_3$OH with respect
to water along with the relative CH$_3$OH/CO abundance ratio,
as observed towards low- (LM), intermediate- (IM), and high-mass (HM) protostars. }
\begin{tabular}{c c c c}
\hline
\hline
Source & X(CO) & X(CH$_3$OH) & X(CH$_3$OH) \\
& \% wrt H$_2$O & \% wrt H$_2$O & \% wrt to CO \\
\hline
\textbf{Low-Mass} & & & \\
L1448 IRS 1 & 45.5 \tablefootmark{a-e} & $<14.9$ \tablefootmark{d} & $<32.7$ \\
L1455 SMM 1 & 5.6 \tablefootmark{a-e} & $<13.5$ \tablefootmark{d} & $<241.1$ \\
RNO 15 & 13.6 \tablefootmark{a-e}& $<5.0 $ \tablefootmark{d} & $<36.8$ \\
IRAS 03254 & 13.6 \tablefootmark{a-e} & $<4.6 $ \tablefootmark{d} & $<33.8$ \\
IRAS 03271 & 8.2 \tablefootmark{a-e} & $<5.6$ \tablefootmark{d} & $<68.3$ \\
B1-a & 13 \tablefootmark{a-e} & $<1.9$ \tablefootmark{d} & $<14.6$ \\
B1-c & 28.6 \tablefootmark{a-e} & $<7.1$ \tablefootmark{d} & $<24.8$ \\
L1489 & 15.2 \tablefootmark{a-e} & $4.9$ \tablefootmark{d} & 32.3 \\
DG Tau B & 20.5 \tablefootmark{a-e} & $<5.7$ \tablefootmark{d} & $<27.8$ \\
IRAS 12553 & 12.6 \tablefootmark{a-e} & $<3.0$ \tablefootmark{d} & $<23.8$ \\
IRAS 13546 & 22.7 \tablefootmark{a-e} & $<3.9$ \tablefootmark{d} & $<17.2$ \\
IRAS 15398 & 40.3 \tablefootmark{a-e} & 10.3 \tablefootmark{d} & 25.6 \\
CRBR 2422.8-3423 & 11.2 \tablefootmark{a-e} & $<9.3$ \tablefootmark{d} & $<83.0$ \\
RNO 91 & 19.0 \tablefootmark{a-e} & $<5.6$ \tablefootmark{d} & $<29.5$ \\
IRAS 23238 & 4.0 \tablefootmark{a-e} & $<3.6$ \tablefootmark{d} & $90.0$ \\
\hline
\textbf{Intermediate-Mass} & & & \\
AFGL989 & 19.0 \tablefootmark{b} & 1.7 \tablefootmark{b} & 8.9 \\
SMM4 & 30.0 \tablefootmark{c} & 28.0 \tablefootmark{c} & 93.3 \\
\hline
\textbf{High-Mass} & & & \\
W33A & 7.4 \tablefootmark{a-e} & 14.7 \tablefootmark{d} & 198.6 \\
GL 2136 & 10.2 \tablefootmark{a-e} & 8.5 \tablefootmark{d} & 83.3 \\
S140 & 16.6 \tablefootmark{a-e} & $<3.0$ \tablefootmark{d} & $<18.1$ \\
NGC 7538 IRS 9 & 16.5 \tablefootmark{a-e} & 7.5 \tablefootmark{d} & 45.5 \\
AFGL2136 & 5.0 \tablefootmark{b} & 5.0 \tablefootmark{b} & 100.0 \\
AFGL7009 & 16.0 \tablefootmark{b} & 33.0 \tablefootmark{b} & 206.3 \\
\hline
\end{tabular}
\label{obs_CO_CH3OH}
\tablebib{$^{(a)}$ \citet{Pontoppidan2003}; $^{(b)}$ \citet{Gibb2004};
$^{(c)}$ \citet{Pontoppidan2004}; $^{(d)}$ \citet{Boogert2008};
$^{(e)}$ \citet{Pontoppidan2008}.}
\end{table}
\subsection{Constraining the activation and diffusion energies}
As a first step, we compared the model predictions with the observations
with the goal to constrain the two microphysics parameters, which have
a high impact on the model predictions (\S 4.6), namely $E_d/E_b$ and $E_a$. Figure
\ref{fig:obs-EaEd} shows the CH$_3$OH/CO abundance ratio as a function
of time obtained assuming reference parameters (and more particularly
a density $10^5$ cm$^{-3}$) and for different
values of the activation energy $E_a$ and diffusion energy $E_d$.
The comparison of the model predictions with the values
observed towards intermediate- and high- mass protostars clearly
excludes an activation energy higher than 1450 K and a
diffusion energy higher than 0.65 times the binding energy $E_b$.
More stringently, if $E_a = 1450$ K, $E_d$/$E_b$ has to be equal to
0.5. The analysis of a similar plot for a density of $10^4$ cm$^{-3}$
gives similar constraints: $E_a$=1450 K implies $E_d$/$E_b$=0.5 and
$E_a$=400 K, $E_d$/$E_b\sim 0.65$. The case $10^6$ cm$^{-3}$ does not
provide more constraints, it is indeed excluded by the observations
(see below).
\begin{figure}[tbh]
\centering
\includegraphics[width=88mm]{Ea_diff_Xmantle_CH3OH_CO_obs.ps}
\caption{Mantle CH$_3$OH/CO abundance ratio versus time. The blue,
green, and red lines refer to $E_d$/$E_b$=0.65, 0.5, and 0.8,
respectively. Solid, dotted, and dashed lines refer to $E_a$=400 K,
1450 K, and 2500 K, respectively. In these computations the density is
$10^5$ cm$^{-3}$. The box with hatching shows the interval of CH$_3$OH/CO
abundance ratio observed towards intermediate- and high- mass
protostars (Table \ref{obs_CO_CH3OH}). Because low-mass protostars
provide only less stringent upper limits, they are not reported in
the plot.}
\label{fig:obs-EaEd}
\end{figure}
\subsection{Constraining the pre-collapse phase duration and density}
Once we limited the range of possible values of $E_a$ and $E_b$
consistent with the observations, we can attempt to constrain the
duration of the pre-collapse phase and the density. To this end,
Figure \ref{fig:mod-obs-dens} shows the curves of the CH$_3$OH/CO
abundance ratio versus time for three different densities ($10^4$,
10$^5$ and $10^6$ cm$^{-3}$) and computed for the two following sets
of $E_a$ and $E_d$ values: 1) $E_a$=400 K and $E_d/E_b$=0.65 (and
the reference parameters) ; 2) $E_a$=1450 K and $E_d/E_b$=0.5 (and
the reference parameters).
\begin{figure}[tbh]
\centering
\includegraphics[width=88mm]{nH_2models_multilayer_Xmantle_CH3OH_CO_obs.ps}
\caption{Mantle CH$_3$OH/CO abundance ratio versus time. The red lines
refer to Model 1 ($E_a$=400 K and $E_d$/$E_b$=0.65), the blue lines
refer to Model 2 ($E_a$=1450 K and $E_d/E_b$=0.5) (see text).
Solid, dotted and dashed lines refer to densities $10^4$,
$10^5$ and $10^6$ cm$^{-3}$ respectively. The grey box with hatching shows
the interval of CH$_3$OH/CO abundance ratio observed towards high-
and intermediate- mass protostars, while the green dashed box shows
the values (upper limits) observed in low-mass protostars (Table
\ref{obs_CO_CH3OH}).}
\label{fig:mod-obs-dens}
\end{figure}
The comparison between the predicted and observed CH$_3$OH/CO
abundance ratio in intermediate- and high-mass protostars suggests
that the bulk of methanol has been formed
when the pre-collapse condensation had densities between $10^4$ and
$10^5$ cm$^{-3}$. Higher densities would produce an insufficiently CH$_3$OH/CO
abundance ratio compared to the values observed in intermediate-
and high- mass protostars. If the densities are around $10^5$
cm$^{-3}$, the duration of the pre-collapse phase must have lasted
at least $10^5$ yr. Lower densities would allow shorter pre-collapse
duration times. But densities higher than $10^5$ cm$^{-3}$ are
not ruled out for the condensations that produce the low-mass
protostars. In that case, the duration of the pre-collapse phase
lasted more than $5\times10^4$ yr.
One important result is that the CH$_3$OH/CO abundance ratio does not
provide an estimate of the duration of the pre-collapse phase, but
only a lower limit to it. It neither provides an estimate of the
density at the time of the collapse but only a range of possible
densities for any given observed CH$_3$OH/CO abundance ratio. Note
that the CH$_3$OH/H$_2$CO abundance ratio does not allow us to arrive at better
constraints, because the theoretical curves are very similar to those of
Fig. \ref{fig:mod-obs-dens}.
\subsection{Chemical differentiation within grain mantles}
As briefly mentioned in \S 2, IR observations suggest a chemical
differentiation of the grain mantles with a polar matrix formed
already at $A_V \sim 3$ mag \citep{Whittet1988}, and a non-polar CO
matrix formed at higher densities \citep{Tielens1991}. In between, CO
can be mixed either with water, methanol, or both at the same time
\citep{Bisschop2007}, as well as mixed with O$_2$, O$_3$, N$_2$, or
CO$_2$ \citep{Pontoppidan2006}. Actually, this observational fact is
one of the motivations of our work. As discussed in \S
\ref{sec:results} and, for example, shown in Fig. \ref{ML_compo}, our
code predicts inhomogeneous grain mantle composition. For the
reference model, for example, CO is mixed with water in the inner
layers while a mixture between CO and methanol is predicted in the
outer layers. Given the lack of a dataset to compare our results with, it is
difficult to carry out quantitative comparisons between observations
and predictions. However, it is worth noticing that the exact grain
mantle composition and stratification depends on the overall evolution
and structure of the pre-collapse condensation, because the chemical
differentiation depends on the density, the temperature and time.
\section{Conclusions and perspectives}\label{sec:concl-persp}
We presented a new model, GRAINOBLE, which computes the chemical
composition of interstellar grain mantles during the cold and dense
phase of the pre-collapse. The model presents two main differences
from most other published codes: 1) it assumes that only the outermost
mantle layer is chemically active, while the mantle bulk is not; 2) it
considers porous grains.
We run GRAINOBLE for a large set of parameters, which are either
unconstrained (density, temperature and grain size of the prestellar
condensation) or poorly known (gas atomic oxygen abundance,
diffusion energy, hydrogenation reactions activation energy, grain
site size) . We obtained a grid of about 18000 models and built
distribution plots of the predicted mantle abundances. This allows us
to study the influence of each of these parameters on the predicted
abundances and, consequently, the robustness of the predictions.
Finally, we compared the GRAINOBLE predictions with those obtained by
the microscopic CTRW model of \citet{Cuppen2009}, obtaining a fair
agreement between the two models. The comparison clearly validates our
treatment and our code (\S \ref{sec:comp-with-prev}).
The most important results of the GRAINOBLE model are the following \\
1) The multilayer treatment shows a differentiation of the species
within the mantle, with the innermost layers being rich in CO, the
intermediate layers in formaldehyde, and the outermost layers in
methanol, reflecting the different formation time of each species in
the mantle. This differentiation will likely lead to a
differentiation in the deuteration of formaldehyde and methanol on the
ices. Because the deuteration increases with the CO depletion in the gas
phase \citep{Roberts2003, Ceccarelli2005}, methanol will likely be
more deuterated than formaldehyde, as indeed observed in
\citet{Parise2006}. A forthcoming paper will focus on the detailed
modeling of the deuteration process. \\
2) The multilayer treatment predicts a relatively high abundance of
radicals trapped in the mantle. {For example, HCO and CH$_3$O show abundances
of $10^{-9}-10^{-7}$ with respect to H nuclei. As suggested by Garrod \& Herbst
(2006), these radicals can react to form COMs when the grain
temperature increases during the formation of the protostar. At this
stage, we cannot say whether the predicted radical abundances are
sufficient to explain the present observations (e.g. Fig. 1), and,
unfortunately, comparisons with previous models are not possible
because of the lack of specific information.
We will explore this aspect, the formation of COMs, in a forthcoming paper. \\
3) The presence of porosity in the grains only moderately influences
the mantle chemical composition, causing an enhanced abundance of
formaldehyde and methanol by less than a factor four.\\
4) The chemical composition of grain mantles strongly depends on the
physical conditions of the prestellar condensation, particularly on
its density as well as its age. Therefore, it will be important to
model the evolution of the condensation for realistic predictions of
the mantle composition. Conversely, the observed mantle composition
can provide valuable information on the past history of protostars,
and will be the focus of a forthcoming paper. Comparisons of the
present predictions with the ices observations (specifically the
CH$_3$OH/CO abundance ratio) suggest that intermediate- and high- mass
protostars evolved from condensations less dense than about $10^5$
cm$^{-3}$, while no stringent constraints can be put on the prestellar
condensations of low-mass protostars.\\
5) The predicted mantle composition critically depends on the values
of the diffusion energy and activation energy of the hydrogenation
reactions. Comparing our model predictions with observations of ices
allows to constrain the values of these two important parameters: the
diffusion to binding energy ratio has to be around 0.5--0.65, and the
activation energy has to be less than 1450 K.\\
6) Other parameters of the GRAINOBLE model, like the gas atomic oxygen
abundance and the site size, do not substantially influence the
predicted mantle abundances.
In summary, GRAINOBLE is a versatile and fast code to model the grain
surface chemistry. Its main improvement with respect to previous
similar models is the treatment of the layer-by-layer chemistry, which
provides a more realistic stratified mantle composition. It predicts
that the mantles are indeed stratified and that radicals are trapped
inside the mantles. Because it is not expensive from the point of view
of computing time, modeling of complex and more realistic cases, like
the evolution of a prestellar core, are now feasible.
\begin{acknowledgements}
The authors would like to thank
O. Biham, P. Caselli, T. Hasegawa, and V. Wakelam for useful exchanges on grain chemistry modelling,
P. Peters and L. Wiesenfeld for discussions about the physical and chemical processes on grain surfaces,
and X. Tielens for helpful discussions on the abundances in the gas phase.
This work has been supported by l\textquoteright Agence Nationale pour la Recherche (ANR), France (project FORCOMS, contracts ANR-08-BLAN-022).
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction} \label{sec:intro}
The {\it Herschel} surveys have allowed clustering studies (Maddox
et al. 2010; Cooray et al. 2010) of sub-millimeter galaxies with a
statistics at least one order of magnitude better than previously
possible (Blain et al. 2004; Scott et al. 2006). These studies have
been complemented by determinations of the angular power spectrum of
the Cosmic Infrared Background (CIB) anisotropies on BLAST (Viero et
al. 2009), {\it Planck} (Planck Collaboration 2011), and {\it
Herschel} (Amblard et al. 2011) maps. Due to the unique power of
sub-millimeter surveys in piercing the distant universe, thanks to
the strongly negative K-correction, the clustering properties
contain signatures of the large scale structure at high redshifts
and can allow us to discriminate between different formation
mechanisms that have been proposed for sub-millimeter galaxies. For
example, merger driven galaxy evolution models, that follow the
evolution of both the disk and the spheroidal components of
galaxies, predict much lower clustering strengths for sub-mm
galaxies (e.g. Almeida et al. 2011; Kim et al. 2011) than models
whereby the star formation is fueled by steady accretion of large
amounts of cold gas (e.g. Dav\'e et al. 2010).
In this paper, building on the work by Negrello et al. (2007), we
investigate the constraints set by mm and sub-mm clustering data on
the physical model worked out by Granato et al. (2001, 2004) and
further elaborated by Lapi et al. (2006) and Mao et al. (2007).
A specific prediction of the model is that high-$z$ massive
proto-spheroidal galaxies dominate the sub-mm counts over a limited
flux density range (cf. Lapi et al. 2011). At $250\,\mu$m the
Euclidean normalized differential counts of these objects peak at
$\approx 30$ mJy; above $\simeq 60\,$mJy and below $\simeq 10\,$mJy
the counts are dominated by $z\,\lower2truept\hbox{${<\atop\hbox{\raise4truept\hbox{$\sim$}}}$}\, 1.5$ quiescent and star-bursting
late-type galaxies, less massive and less clustered than the
high-$z$ proto-spheroidal galaxies. The flux density range where
proto-spheroidal galaxies dominate broadens and the peak shifts to
brighter flux densities with increasing (sub-)mm wavelength.
Therefore, in this scenario, the expected clustering strengths
depend on the flux density range that is being probed and on
wavelength.
Several other analyses of data on the angular correlation function
of (sub-)mm sources and of the power spectrum of the CIB
anisotropies have been carried out. They however use
phenomenological parameterized models for the evolution of
extragalactic sources (Hall et al. 2010; Planck Collaboration 2011;
Millea et al. 2011; P\'enin et al. 2011) or even of the clustering
power (Addison et al. 2011). Also data at different wavelengths are
usually fitted separately (Planck Collaboration 2011; Amblard et al.
2011). On the contrary, the present analysis relies on a physical
model for the evolution of proto-spheroidal galaxies (although the
treatment of spiral and starburst galaxies is phenomenological) and
aims at accounting simultaneously for clustering data over a broad
range of wavelengths, from $250\,\mu$m to a few mm.
It should be noted, however, that the physical model is exploited
only to compute the cumulative flux function that weights the
redshift-dependent spatial power spectrum in the Limber
approximation for the angular power spectrum. The Halo Occupation
Distribution (HOD), which is a statistical description of how dark
matter halos are populated with galaxies, is dealt with in a
simplified manner, without including the relationship between
luminosity and halo mass. This is the standard practice, justified
by the complexity of a thorough treatment that does not appear to be
required by existing data. In the Granato et al. (2004) model the
star-formation rate is related to the halo mass, {\it to the
virialization redshift and to the age of the galaxy}. Including
these additional ingredients in the analysis is impractical at the
present stage. A pioneering model that explicitly includes a
relationship between infrared luminosity and halo mass has been
presented by Shang et al. (2011).
The plan of the paper is the following. In \S\,\ref{sec:model} we
present a short overview of the evolutionary model for the relevant
galaxy populations. In \S\,\ref{sect:hm} we describe the halo model
formalism used to compute the contributions to the power spectrum of
Cosmic Infrared Background (CIB) anisotropies and to the angular
correlation function of detected galaxies (\S\,\ref{sect:ps}). Our
main results are presented in \S\,\ref{sect:results} and our main
conclusions are summarized in \S\,\ref{sect:conclusions}.
We adopt a standard flat $\Lambda$CDM cosmology with
$h=H_0/100\,\hbox{km}\,\hbox{s}^{-1}\,\hbox{Mpc}^{-1}=0.70$ and a
local matter density $\Omega_{{\rm m}0}=0.27$.
\section{Overview of the model} \label{sec:model}
The sub-millimeter extragalactic sources are a mixed bag of various
populations of dusty galaxies and of flat-spectrum radio sources
(see, e.g., Lapi et al. 2011).
Our model interprets powerful high-$z$ sub-mm galaxies as massive
proto-spheroidal galaxies in the process of forming most of their
stellar mass (see also Blain et al. 2004; Narayanan et al. 2010;
Dav\'e et al. 2010). It hinges upon high resolution numerical
simulations showing that dark matter halos form in two stages (Zhao
et al. 2003; Wang et al. 2011; Lapi \& Cavaliere 2011). An early
fast collapse of the halo bulk, including a few major merger events,
reshuffles the gravitational potential and causes the dark matter
and the stellar component to undergo (incomplete) dynamical
relaxation. A slow growth of the halo outskirts in the form of many
minor mergers and diffuse accretion follows; this second stage has
little effect on the inner potential well where the visible galaxy
resides.
The star formation is triggered by the fast collapse/merger phase of
the halo and is controlled by self-regulated baryonic processes. It
is driven by the rapid cooling of the gas within a region of $\simeq
70(M_h/10^{13}\,M_\odot)^{1/3}[(1+z)/3]^{-1}\,$kpc, where $M_h$ is
the halo mass, is regulated by the energy feedback from supernovae
(SNe) and Active Galactic Nuclei (AGNs), is very soon obscured by
dust and is stopped by quasar feedback. The AGN feedback is
relevant especially in the most massive galaxies and is responsible
for their shorter duration ($5-7\times 10^8\,$yr) of the active
star-forming phase. In less massive proto-spheroidal galaxies the
star formation rate is mostly regulated by SN feedback and continues
for a few Gyr.
Since spheroidal galaxies are observed to be
in passive evolution at $z\la 1-1.5$ (e.g., Renzini 2006), they are
visible at sub-mm wavelength only at high redshifts. Lapi et al.
(2011) have shown that the Granato et al. (2004) model, as further
elaborated by Lapi et al. (2006), provides a reasonably good fit to
the observed counts from $250\,\mu$m to $\simeq 1\,$mm as well as to
the luminosity functions in the range $z=1-4$ and to the redshift
distributions at $z>1$ estimated from {\it Herschel}-ATLAS (Eales et
al. 2010) data.
The fit was obtained using of a single SED (that of the well studied
$z=2.3$ strongly lensed galaxy SMM~J2135-0102, ``The Cosmic
Eyelash''; Ivison et al. 2010, Swinbank et al. 2010) for the whole
population of proto-spheroidal galaxies. This is obviously an
oversimplification and indeed the Lapi et al. (2011) counts are
somewhat high at mm wavelengths, especially at relatively bright
flux densities. As a consequence, the model overestimates the
Poisson (shot-noise) contribution to the power spectrum of intensity
fluctuations since such contribution is directly related to the
source counts [see eq.~(\ref{eq:shot})]. Consistency with the
shot-noise levels estimated by Planck Collaboration (2011) at 353
GHz and measured by Hall et al. (2010), Dunkley et al. (2011) at 220
and 150 GHz is recovered scaling down the cumulative flux function
$dS/dz$ [see eq.~(\ref{eq:dSdz})] of proto-spheroidal galaxies by a
factor of 0.81, 0.71, and 0.55 at 353, 217, and 150 GHz
($850\,\mu$m, 1.38\,mm, 2\,mm), respectively. No correction was
applied at higher frequencies. In practice, we use the determination
of the shot noise amplitude to recalibrate the function $dS/dz$ to
be used to compute the clustering power spectrum, which is measured
independently. This correction mimics the effect of adopting a SED
decreasing with increasing wavelength beyond the peak a bit more
steeply than the one adopted by Lapi et al. (2011).
As suggested in the latter paper, the overestimate of mm-wave counts
may be cured if higher-$z$ galaxies, that yield larger and larger
contributions to the bright counts at increasing mm wavelengths,
have SEDs slightly hotter than SMM J2135-0102 and closer to that of
G15.141 (Cox et al. 2011; see Fig. 2 of Lapi et al. 2011). We have
checked that indeed a good fit of the counts at all the frequencies
considered here is obtained using the SMM J2135-0102 SED for
galaxies at $z<3.5$ and the SED of G15.141 at higher z. However the
match of the frequency spectrum of the shot-noise amplitude also
improves but not enough to reach consistency with observational
estimates at the longer wavelengths. Since the shot noise amplitude
can be computed directly from the counts, this suggests that there
may be some small, but non-negligible, offsets between the
calibration of point source flux densities and that of the diffuse
background. This is not surprising since, in addition to the
possibility of an imperfect photometric calibration, at mm
wavelengths the recovery of the contribution of dusty galaxies to
the power spectrum requires a delicate subtraction of the other
components (Cosmic Microwave Background, cirrus emission,
fluctuations due to radio sources). A rescaling to match the shot
noise spectrum seems to be the only practical way for correcting for
these offsets. Since the modification is only significant at $\ge
850\,\mu$m (in the observer frame), i.e. well beyond the peak for
most sources, the impact on the bolometric luminosity, which is
related to the halo mass, is minor. For galaxies at redshifts up to
$z=3.5$, accounting for essentially all the signal, the bolometric
luminosity varies by $\le 2\%$. For comparison, the coefficient of
the relationship between the star formation rate (SFR; given by the
model) and the bolometric luminosity has an uncertainty of $\sim
30\%$ (Kennicutt 1998).
The Granato et al. (2004) model is meant to take into account the
star formation occurring within galactic dark-matter halos
virialized at $z_{\rm vir} \,\lower2truept\hbox{${> \atop\hbox{\raise4truept\hbox{$\sim$}}}$}\, 1.5$ and bigger than $M_{\rm vir}
\simeq 10^{11.2} M_\odot$, which are, crudely, associated to massive
spheroidal galaxies. We envisage disk (and irregular) galaxies as
associated primarily to halos virializing at $z_{\rm vir} \,\lower2truept\hbox{${<\atop\hbox{\raise4truept\hbox{$\sim$}}}$}\,
1.5$, which have incorporated, through merging processes, a large
fraction of halos less massive than $10^{11.2}\,M_\odot$ virializing
at earlier times, which may become the bulges of late type galaxies.
The model, however, does not follow the formation and evolution of
disk and bulge components of galaxies. For spiral and starburst
galaxies we adopt the phenomenological model described by Negrello
et al. (2007). On the other hand, as shown in the following, these
galaxies are essentially non influential for the purposes of the
present paper in the considered frequency range: proto-spheroids
dominate the contributions both to the power spectrum of
fluctuations and to the angular correlation function of detected
sources.
Because of the strong dilution due to their very broad luminosity
function, the contribution of radio sources to the clustering power
spectrum can be safely neglected in the wavelength range considered
here. Their contribution to Poisson fluctuations was computed using
the De Zotti et al. (2005) model.
\section{Halo Model Formalism}\label{sect:hm}
To compare the clustering properties expected from our model with
observational data we adopt the halo model formalism (Cooray \&
Sheth 2002). The power spectrum of the galaxy distribution is
parameterized as the sum of the 1-halo term, that dominates on small
scales and depends on the distribution of galaxies within the same
halo, and the 2-halo term, that dominates on large scales and is
related to correlations among different halos:
\begin{eqnarray}
P_{\rm gal}(k,z)\!\!\!\!\!\!&=&\!\!\!\!\!\!P^{\rm 1h}_{\rm
gal}(k,z)+P^{\rm 2h}_{\rm gal}(k,z),\hfill \\
P^{\rm 1h}_{\rm gal}(k,z)\!\!\!\!\!\!&=&\!\!\!\!\!\!\int_{M}\!\!\!\!dM\frac{dn}
{dM}\frac{\langle{N_{\rm gal}(N_{\rm gal}-1)}\rangle}{\bar{n}^2_{\rm gal}}|u_{\rm gal}(k,M)|^s,\label{eq:1halo}\\
P^{\rm 2h}_{\rm gal}(k,z)\!\!\!\!\!\!&=&\!\!\!\!\!\!P_{\rm
lin}(k,z)\!\! \left[\!\int_M\!\!\!\!\!
dM\!\frac{dn}{dM}\frac{\langle{N_{\rm gal}}\rangle} {\bar{n}_{\rm
gal}}b(M,\!z)u_{\rm gal}(k,\!M)\right]^2\!\!\!,\label{eq:2halo}
\end{eqnarray}
where $dn/dM$ is the halo mass function (Sheth \& Tormen 1999) and
the linear matter power spectrum, $P_{\rm lin}(k,z)$, has been
computed using the CAMB code\footnote{http://camb.info/} (Lewis,
Challinor \& Lasenby 2000). Here, $u_{\rm gal}(k,M)$ denotes the
Fourier transform of the mass density profile of the galaxy
distribution within the dark matter halo, that we assume to be
approximately the same as that of the dark matter, i.e. we take
$u_{\rm gal}(k,M)\simeq u_{\rm dm}(k,M)$.
${\langle{N_{\rm gal}}\rangle}$ is the mean number of galaxies in a
halo of mass $M$, subdivided in ``central'' and ``satellite''
galaxies (${\langle{N_{\rm gal}}\rangle}= {\langle{N_{\rm
cen}}\rangle}+{\langle{N_{\rm sat}}\rangle}$), while $\bar{n}_{\rm
gal}$ is the mean number density of galaxies:
\begin{equation}
\bar{n}_{\rm gal}=\int_{M}dM\frac{dn}{dM}{\langle{N_{\rm gal}}\rangle}.
\end{equation}
We model the HOD using a central-satellite formalism (see, e.g., Zheng et al. 2005): this
assumes that the first galaxy to be hosted by a halo lies at its
center, while any remaining galaxies are classified as satellites
and are distributed in proportion to
the halo mass profile.
Following Tinker \& Wetzel (2010), the mean occupation functions of
central and satellite galaxies are parameterized as:
\begin{eqnarray}
{\langle{N_{\rm cen}}\rangle}\!\!\!\!\!\!&=&\!\!\!\!\!\!\frac{1}{2}
\left[1+{\rm erf}\left({\frac{\log_{10}(M/M_{\rm min})}{\sigma(\log_{10}M)}}\right)\right],\\
{\langle{N_{\rm
sat}}\rangle}\!\!\!\!\!\!&=&\!\!\!\!\!\!\frac{1}{2}\left[1+{\rm
erf}\left({\frac{\log_{10}(M/2M_{\rm
min})}{\sigma(\log_{10}M)}}\right)\right]\left(\frac{M}{M_{\rm
sat}}\right)^{\alpha_{\rm sat}},
\end{eqnarray}
where $M_{\rm min}$, $\alpha_{\rm sat}$, $M_{\rm sat}$, and
$\sigma(\log_{10}M)$ are free parameters assumed to be redshift
independent. In this formalism halos below $M_{\rm min}$ do not
contain galaxies while halos above this threshold contain a central
galaxy plus a number of satellite galaxies with a power-law mass
function with slope $\alpha_{\rm sat}$.
The mean mass density profile of halos of mass $M$ is (Navarro,
Frenk, \& White 1996):
\begin{equation}
\rho(r)=\frac{\rho_{\rm s}}{(r/r_{\rm s})(1+r/r_{\rm s})^{2}},
\end{equation}\label{eq:NFW}
\begin{equation}
M=4\pi\rho_{\rm s}r^3_{\rm s}\left[\log(1+c)-\frac{c}{1+c}\right],
\end{equation}
with $c=r_{\rm vir}/r_{\rm s}$. The normalized Fourier transform of
this profile is:
\begin{eqnarray}
\lefteqn{u_{\rm dm}(k,M)=\frac{4\pi\rho_{\rm s}r^3_{\rm
s}}{M}\left\{\sin(kr_{\rm s})\left[Si([1+c]kr_{\rm s})-Si(kr_{\rm
s})\right]\frac{}{}\right.}\nonumber \\
& +&\!\!\!\!\!\! \left.\cos(kr_{\rm s})\left[Ci([1+c]kr_{\rm
s})-Ci(kr_{\rm s})\right]-\frac{\sin(ckr_{\rm s})}{(1+c)kr_{\rm
s}}\right\},
\end{eqnarray}
where $Si$ and $Ci$ are the sine and cosine integrals, respectively:
\begin{equation}
Si(x)=\int^x_0\frac{\sin(t)}{t}dt,~~~Ci(x)=-\int^\infty_x\frac{\cos(t)}{t}dt.
\end{equation}
Following Bullock et al. (2001), we approximate the dependence of
the concentration $c$ on $M$ and $z$ as
\begin{equation}
c(M,z)=\frac{9}{1+z}\left(\frac{M}{M_\ast}\right)^{-0.13}
\end{equation}
where $M_\ast(z)$ is the characteristic mass scale at which
$\nu(M,z)=1$; $M_\ast(z=0)\simeq5\times10^{12}\,h^{-1}\,M_\odot$.
In the 1-halo term [eq.~(\ref{eq:1halo})] we set $s=2$, in analogy
with the corresponding term for the dark matter power spectrum, if
${\langle{N_{\rm gal}(N_{\rm gal}-1)}\rangle} > 1$. Otherwise we set
$s=1$ since if the halo contains only one galaxy, it will sit at the
center. Taking into account that ${\langle{N_{\rm
gal}(N_{\rm gal}-1)}\rangle} \simeq 2{\langle{N_{\rm
cen}\rangle\langle N_{\rm sat}}\rangle}+{\langle{N_{\rm
sat}\rangle}}^2$ and that only the galaxies that are not at the
center get factors of $u_{\rm gal}(k,M)\simeq u_{\rm dm}(k,M)$ we
have:
\begin{eqnarray}
\lefteqn{P^{\rm 1h}_{\rm gal}(k,z)=\frac{1}{\bar{n}_{\rm
gal}^2}\int_{M}dM\frac{dn}{dM} \cdot} \nonumber \\
&\cdot&
\left[2{\langle{N_{\rm cen} \rangle\langle N_{\rm
sat}}\rangle}u_{\rm dm}(k,M)+{\langle{N_{\rm sat}\rangle}}^2u^2_{\rm
dm}(k,M)\right],
\end{eqnarray}
\begin{eqnarray}
\lefteqn{P^{\rm 2h}_{\rm gal}(k,z)=P_{\rm lin}(k,z) \cdot} \nonumber
\\
&\cdot& \left[{\int_M dM\frac{dn}{dM}\frac{\langle{N_{\rm
gal}}\rangle}{\bar{n}_{\rm gal}}b(M,z)u_{\rm dm}(k,M)}\right]^2,
\end{eqnarray}
with $(dn/dM)dM =f(\nu)(\rho_{\rm m}/M)d\nu$,
\begin{equation}
\bar{n}_{\rm gal}=\int_{M}dM\frac{dn}{dM}{\langle{N_{\rm
gal}}\rangle}=\int_{\nu}d\nu{f(\nu)}\left(\frac{\rho_{\rm
m}}{M}\right){\langle{N_{\rm gal}}\rangle},
\end{equation}
and $\rho_{\rm m}/M=3/(4\pi R^3)$. On large scales, where the 2-halo
term dominates, $u_{\rm dm}(k,M)\rightarrow1$ and $P^{\rm 2h}_{\rm
dm}(k,z)\simeq b_{\rm gal}^2 P_{\rm lin}(k,z)$ with:
\begin{equation}
b_{\rm gal}(z)=\int_{\nu}d\nu{f(\nu)}\left(\frac{\rho_{\rm
m}}{M}\right)b(M,z)\frac{\langle{N_{\rm gal}\rangle}}{\bar{n}_{\rm
gal}}.\label{eq:galbias}
\end{equation}
We also define the effective large-scale bias, $b_{\rm eff}(z)$, as
\begin{equation}
b_{\rm eff}(z) = \int_{M}dM\,\frac{dn}{dM}\frac{\langle{N_{\rm gal}}\rangle}{\bar{n}_{\rm gal}}b(M,z),
\label{eq:beff}
\end{equation}
and the effective mass of the halo, $M_{\rm eff}$,
\begin{equation}
M_{\rm eff}(z) = \int_{M} dM\,\frac{dn}{dM} M \frac{\langle{N_{\rm gal}}\rangle}{\bar{n}_{\rm gal}}.
\label{eq:Meff}
\end{equation}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.28]{1200GHz.eps}
\includegraphics[scale=0.28]{857GHz.eps}
\includegraphics[scale=0.28]{545GHz.eps}
\includegraphics[scale=0.28]{353GHz.eps}
\caption{CIB angular power spectra $P(k_{\theta})$ at sub-mm
wavelengths. Data from Planck Collaboration (2011) and {\it
Herschel}/HerMES (Amblard et al. 2011). For the {\it
Herschel}/HerMES data at 350 and $500\,\mu$m we have adopted the
values corrected by Planck Collaboration (2011). At $250\,\mu$m we
have used the values given by Amblard et al. (2011), that may be
underestimated because of an over-subtraction of the cirrus
contamination and a slight overestimate of the effective beam area.
The conversion from the multipole number $\ell$ used by Planck
Collaboration (2011) and the wavenumber $k\,(\hbox{arcmin}^{-1})$ is
$k=\ell/(2\times 180\times 60)$. The lines show the contributions of
the 1-halo and 2-halo terms for the two populations considered here
[spiral and starburst (SS), and proto-spheroidal (PS) galaxies]. The
magenta horizonal lines denote the shot noise level.
}\label{fig:pk_planck}
\end{center}
\end{figure*}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.20]{217GHz.eps}
\includegraphics[scale=0.20]{150GHz.eps}
\includegraphics[scale=0.20]{150x220GHz.eps}
\caption{CIB angular power spectra $P(k_{\theta})$ at mm wavelengths
and $150\times 220\,$GHz cross spectrum. Data from Planck
Collaboration (2011) and SPT (Hall et al. 2010; Shirokoff et al.
2011). The ACT data (Dunkley et al. 2011; Das et al. 2011) are in
good agreement with the SPT ones and are not plotted to avoid
over-crowding the figure. The lines have the same meaning as in
Fig.~\protect\ref{fig:pk_planck}. }\label{fig:pk_SPT}
\end{center}
\end{figure*}
\section{Angular Power Spectrum of Intensity Fluctuations}\label{sect:ps}
The angular power spectrum $P(k_\theta)$ of intensity fluctuations
due to clustering of sources fainter than some flux density limit
$S_{\rm lim}$ is a projection of the spatial power spectrum of such
sources at different redshifts $z$, $P_{\rm gal}(k,z)$. In the Limber
approximation, valid if the angular scale is not too large (i.e.
$2\pi k_\theta\geq 10$), the relation between $P_{\rm gal}(k,z)$ and
$P(k_\theta)$ is:
\begin{equation}
P(k_\theta)=\!\!\int^{z_{\rm max}}_{z_{\rm min}}\!\!\!\!\!\!\!\!\!\!\!\!dz\,P_{\rm gal}\!\left(\!k=\frac{2\pi
k_\theta+1/2}{\chi(z)},z\right)\left(\frac{dS}{dz}(z)\right)^2\!\!\frac{dz}{dV_{\rm
c}},\label{eq:cl}
\end{equation}
where $dS/dz$ is the redshift distribution of the cumulative flux of
sources with $S\le S_{\rm lim}$
\begin{equation}
{dS\over dz}= \int_0^{ S_{\rm lim}}d\log_{10}(S)\, S\, \phi[L(S,z),z]\,{dV_{\rm c}\over dz},\label{eq:dSdz}
\end{equation}
$\phi(L,z)$ is the epoch-dependent comoving luminosity function per
unit interval of $\log_{10}(L)$, and $dV_{\rm c}$ is the comoving
volume element, $dV_{\rm c}=\chi^2d\chi$, $\chi(z)$ being the
comoving radial distance:
\begin{equation}
\chi(z)=\frac{c}{H_0}\int^z_0\frac{dz'}{\sqrt{\Omega_{\rm m0}(1+z')^3+(1-\Omega_{\rm m0})}}.
\end{equation}
Poisson fluctuations add a white noise contribution to the power
spectrum of fluctuations:
\begin{equation}
P_{\rm shot}=\int^{ S_{\rm lim}}_0 {\frac{dN}{d\log_{10}S}}\,S^2\,{d\log_{10}S},\label{eq:shot}
\end{equation}
with
\begin{equation}
{dN\over d\log_{10}S} = \int\, dz\, \phi[L(S,z),z]\,{dV_{\rm c}\over dz}.
\end{equation}
We have computed the functions $dS/dz$ for each galaxy population
using the the cosmological model specified in \S\,\ref{sec:intro}
and the evolutionary models briefly described in
\S\,\ref{sec:model}. As mentioned in \S\,\ref{sec:model}, the
functions $dS/dz$ for proto-spheroidal galaxies at frequencies $\le
353\,$GHz ($\lambda \ge 850\,\mu$m) have been scaled down by
constant factors to comply with the measurements or the best
estimates of the shot-noise levels. At higher frequencies our model
accurately fits the observed source counts and therefore provides
directly reliable estimates of the shot noise level.
We have chosen to deal with the shot noise and the clustering
contributions to the power spectrum of CIB fluctuations
independently of each other because the former are independent of
the parameters describing the clustering and are strongly
constrained by the available source counts. Moreover, when only
relatively low resolution data are available, as is the case for
{\it Planck}, there is a degeneracy between the shot-noise and the
1-halo clustering term. As clearly highlighted by Planck
Collaboration (2011), an unsupervised least-square fit of the full
CIB power spectrum measured by {\it Planck}, taking the shot-noise
amplitude as a free parameter, leads to fits of similar quality with
and without a substantial contribution from the 1-halo term. But
fits with a low contribution from the 1-halo term imply shot noise
amplitudes far in excess of those estimated from the source counts.
The higher resolution of {\it Herschel}, SPT and ACT data breaks the
degeneracy at $\nu \ge 600\,$GHz and at $\nu \le 220\,$GHz,
respectively, allowing a direct estimate of the shot-noise
amplitude.
As for the halo model, we have considered two distinct populations,
i.e. proto-spheroidal galaxies and late-type galaxies, both
quiescent and starbursting. Taking into account the constraints on
clustering of late-type galaxies coming from IRAS data (Mann et al.
1995; Hawkins et al. 2001) we find that the contribution of these
sources is always sub-dominant and, correspondingly, their halo
model parameters are very poorly constrained. Moreover the values of
$M_{\rm sat}$ and $\sigma(\log_{10}M)$ are poorly constrained also
for proto-spheroidal galaxies (Planck Collaboration 2011). We have
therefore fixed $M_{\rm sat}=20M_{\rm min}$ and
$\sigma(\log_{10}M)=0.6$ [within the ranges found by Tinker \&
Wetzel (2010) from clustering studies of optical galaxies] for both
populations, and $M_{\rm min, late-type} = 10^{11}\,M_\odot$ and
$\alpha_{\rm sat, late-type} = 1$. We are then left with only 2 free
parameters, i.e. $M_{\rm min}$ and $\alpha_{\rm sat}$, for
proto-spheroidal galaxies.
The angular power spectra of CIB anisotropies at 217, 353, 545, and
857 GHz on the multipole range $200 \le \ell \le 2000$ have been
determined by Planck Collaboration (2011) using {\it Planck} maps of
six regions of low Galactic dust emission with a total area of
$140\,\hbox{deg}^2$. In the same paper, the power spectrum
measurements by Amblard et al. (2011), using {\it Herschel}/SPIRE
data at 250, 350, and $500\,\mu$m and extending down to sub-arcmin
angular scales, i.e. up to $\ell \sim 2\times 10^4$, were
re-analyzed. It was found that Amblard et al. (2011) overestimated
the correction for contamination by Galactic cirrus. Moreover, the
diffuse-emission calibration of SPIRE data was improved using the
more accurate {\it Planck}/HFI calibration. We have used the Amblard
et al. (2011) data as corrected by Planck Collaboration (2011) at
350 and $500\,\mu$m. No correction could be applied at $250\,\mu$m
so that the data points at this wavelength could be underestimated.
Power spectrum measurements at mm wavelengths (around 150 and 220
GHz) have been obtained with the SPT and the ACT (Hall et al. 2010;
Dunkley et al. 2011; Shirokoff et al. 2011; Das et al. 2011). The
subtraction of the other components (CMB, Sunyaev-Zeldovich effect,
radio sources) has been done using the best fit values given in the
papers. Note that the units quoted as $\mu\hbox{K}^2$ are actually
$\mu\hbox{K}^2\times$sr. The conversion factor from these units to
$\hbox{Jy}^2/$sr is $\simeq
[24.8(x^2/\sinh(x/2))]^2/[2\ell(\ell+1)]$. The factor
$[24.8(x^2/\sinh(x/2))]^2$ is $\simeq 1.55\times 10^5$ at 150 GHz
and $\simeq 2.34\times 10^5$ at 217 GHz.
The angular correlation function $w(\theta)$ for a single source
population writes, in terms of the 2D power spectrum $P(k_\theta)$:
\begin{equation}
w(\theta) = 2\pi \int_0^\infty k_\theta P(k_\theta) J_0(2\pi
k_\theta \theta) dk_\theta, \label{eq:wtheta}
\end{equation}
where $J_0$ is the Bessel function of order 0.
Here we have two sub-populations, proto-spheroidal and
spirals$+$starburst galaxies, with different clustering properties.
If their cross-correlations can be ignored the signal for the whole
is given by (Wilman et al. 2003):
\begin{equation}
w_{\rm tot}(\theta)=f^2_{\rm PS}w_{\rm PS}(\theta)+f^2_{\rm SS}w_{\rm
SS}(\theta), \label{eq:wtheta_tot}
\end{equation}
where $f_{\rm PS}$ and $f_{\rm SS}$ are the fractional contributions
of proto-spheroidal and spirals$+$starburst galaxies, respectively,
to the total counts:
\begin{equation}
f_{\rm PS/SS}=\frac{\int dz\mathcal{N}_{\rm PS/SS}(z)}{\int
dz\mathcal{N}_{\rm tot}(z)}~,
\end{equation}
$\mathcal{N}$ being the redshift distribution. Ignoring the
cross-correlations between the two source populations is justified
because of the widely different redshift distributions implied by
the adopted evolutionary model: as mentioned in \S\,\ref{sec:model},
proto-spheroidal galaxies are associated to galactic-size halos
virialized at $z_{\rm vir} \,\lower2truept\hbox{${> \atop\hbox{\raise4truept\hbox{$\sim$}}}$}\, 1.5$ while disk (and
irregular/starburst) galaxies are associated primarily to halos
virializing at $z_{\rm vir} \,\lower2truept\hbox{${<\atop\hbox{\raise4truept\hbox{$\sim$}}}$}\, 1.5$.
The spatial correlation function $\xi(r,z)$ is the Fourier
anti-transform of the 3D power spectrum:
\begin{equation}
\xi(r,z)= {1\over 2 \pi^2} \int_0^\infty k^2\, P_{\rm
gal}(k)\,\left(\sin(kr)\over kr \right)\, dk .
\end{equation}
The clustering radius $r_0(z)$ is defined by $\xi(r_0,z)=1$.
\begin{figure}
\begin{center}
\includegraphics[scale=0.28]{dSdz.eps}
\caption{Redshift evolution of the galaxy intensities at 250, 350,
and $500\,\mu$m yielded by the model compared with the
observation-based estimates by Amblard et al. (2011). The different
line styles correspond to the different sub-populations, as
specified in the inset. Some caution is needed in interpreting these
data, in view of the problems pointed out by Planck Collaboration
(2011; see text).}\label{fig:dSdz}
\end{center}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[scale=0.2]{wtheta_250
\includegraphics[scale=0.2]{wtheta_350
\includegraphics[scale=0.2]{wtheta_500
\caption{Angular correlation function of sub-mm galaxies. Data from
Cooray et al. (2010) and Maddox et al. (2010). In the left-hand
panel the dashed and dash-dotted lines show the contributions of PS
and SS populations, while the solid line denotes the total angular
correlation function. In other two panels, the dashed and solid
lines show the global model correlation functions for the limiting
flux densities adopted in the analyses by Maddox et al. and Cooray
et al., respectively. The horizontal dotted line shows the zero
level. \label{fig:wtheta}}
\end{center}
\end{figure*}
\section{Results}\label{sect:results}
Figures~\ref{fig:pk_planck} and \ref{fig:pk_SPT} compare the best
fit model power spectrum with {\it Planck}, {\it Herschel}, and SPT
data in the wavelength range $250\,\mu$m$-2\,$mm. The agreement is
generally good except at $250\,\mu$m where the model is consistently
above the data points by Amblard et al. (2011) which, however, could
be underestimated (see \S\,\ref{sect:ps}). As mentioned above, we
have only two free parameters, i.e. the minimum mass and the
power-law index of the mean occupation function of satellite
galaxies of proto-spheroidal galaxies. The constraints we obtain are
$\log(M_{\rm min}/M_\odot) = 12.24 \pm 0.06$ and $\alpha_{\rm sat} =
1.81 \pm 0.04$ ($1\,\sigma$). The nominal errors on each parameter
have been computed marginalizing on the other and correspond to
$\Delta \chi^2=1$. We caution that the true uncertainties are likely
substantially higher than the nominal values, both because the model
relies on simplifying assumptions that may make it too rigid and
because of possible systematics affecting the data.
The fact that the same values of these parameters account for the
clustering data from 2 mm to $250\,\mu$m confirms the conclusion by
Planck Collaboration (2011) that CIB fluctuations over this
wavelength range are dominated by a single sub-population of dusty
galaxies. According to our model, this sub-population is made of
proto-spheroidal galaxies making most of their stars at $z>1$. We
find that only at $250\,\mu$m other dusty galaxy populations, normal
disk and starburst galaxies, make a significant, but still
sub-dominant contribution to the clustering power spectrum. As shown
by Lapi et al. (2011), according to our model, proto-spheroidal
galaxies also account for the bulk of the CIB intensity in this
wavelength range, consistent with the finding by Planck
Collaboration (2011) that the CIB anisotropies have the same
frequency spectrum as the CIB intensity.
Our estimate of the minimum mass is higher than, but consistent,
within the errors, with those found by Amblard et al. (2011)
considering a single galaxy population and 5 free parameters per
frequency (but one of the parameters is unconstrained by the data
within the prior range): $\log(M_{\rm min}/M_\odot) =
11.1^{+1.0}_{-0.6}$ at $250\,\mu$m, $\log(M_{\rm min}/M_\odot) =
11.5^{+0.7}_{-0.2}$ at $350\,\mu$m, and $\log(M_{\rm min}/M_\odot) =
11.8^{+0.4}_{-0.3}$ at $500\,\mu$m. Our value of $\alpha_{\rm sat}$
is also consistent with those by Amblard et al.: $\alpha_{\rm sat} =
1.6^{+0.1}_{-0.2}$ at $250\,\mu$m, $\alpha_{\rm sat} =
1.8^{+0.1}_{-0.7}$ at $350\,\mu$m and $500\,\mu$m. In Planck
Collaboration (2011) two or three free parameters {\it per
frequency} were used; the derived minimum masses are in the range
$\log(M_{\rm min}/M_\odot) = 11.8-12.5$.
There is however an interesting difference with Planck Collaboration
(2011), due to the different redshift distributions of sources. The
crossover between the 1-halo and the 2-halo term occurs, according
to the model by Planck Collaboration (2011), at multipole numbers
ranging from $\ell \simeq 800$ at 857 GHz ($350\,\mu$m) to $\ell
\simeq 1200$ at 217 GHz ($1.38\,$mm) corresponding to angular scales
ranging from $\theta \simeq 180\times 60/\ell \simeq 13.5'$ at 857
GHz to $9'$ at 217 GHz. According to the B\'ethermin et al. (2011)
model used in that paper, the contribution to the CIB intensity at
857 GHz peaks at $z\simeq 1$ where the angular scale of $13.5'$
corresponds to a physical linear scale $L\simeq 6.5\,$Mpc; at 217
GHz the bulk of the CIB contribution comes from $z>2$ where an
angular scale of $9'$ corresponds to a physical linear scale
$L\simeq 4.5\,$Mpc. The non-linear masses corresponding to an
overdensity $\Delta_c=1.68$ on these scales are $M_{\rm nl}(z=1;
L=6.5\hbox{Mpc})\simeq 7\times 10^{13}\,M_\odot$ and $M_{\rm
nl}(z=2; L=4.5\hbox{Mpc})\simeq 8\times 10^{13}\,M_\odot$,
respectively. For comparison, the characteristic non-linear masses
computed from $\sigma(M_\ast,z)=1.68$ ($\sigma(M_\ast,z)$ being the
rms overdensity) are $M_\ast(z=1)=2 \times 10^{11}\, M_\odot$ and
$M_\ast(z=2)=7.3 \times 10^{9}\, M_\odot$. This suggests that
structures going non-linear on the considered scales are extremely
rare at the corresponding redshifts. This potential difficulty is
eased in our model because the crossover scales are lower by almost
a factor of 2. As shown by Figs.~\ref{fig:pk_planck} and
\ref{fig:pk_SPT}, the 1-halo/2-halo crossover occurs at $\ell \simeq
1450$ at 857 GHz and at $\ell \simeq 2100$ at 217 GHz, corresponding
to angular scales of $7.4'$ and $5.1'$, respectively.
Our value of $M_{\rm min}$ implies an effective halo mass
[eq.~(\ref{eq:Meff})] at $z\simeq 2$ of proto-spheroidal galaxies,
making up most of the CIB, $M_{\rm eff} \simeq 5\times
10^{12}\,M_\odot$, close to the estimated halo mass of the most
effective star formers in the universe. Tacconi et al. (2008)
estimated their mean comoving density at $z\sim 2$ to be $\sim
2\times 10^{-4}\,\hbox{Mpc}^{-3}$. For the standard $\Lambda$CDM
cosmology this implies that they are hosted by dark matter halos of
$\sim 3.5\times 10^{12}\,M_\odot$ (Dekel et al. 2009).
Figure~\ref{fig:dSdz} compares the flux density coming from
different redshifts, $dS/dz$ [eq.~(\ref{eq:dSdz})], predicted by the
model at the SPIRE wavelengths with the best fit estimates by
Amblard et al. (2011). Planck Collaboration (2011) give (their Table
7) the best fit values of the redshift-independent volume
emissivity, $j_{\rm eff}$, for $z>3.5$ [their eq.~(43)]. The values
of $j_{\rm eff}$ given by our model (52, 175, 265, and
$205\,\hbox{Jy}\,\hbox{sr}^{-1}\,\hbox{Mpc}^{-1}$ at 217, 353, 545,
and 857\,GHz, respectively) are consistent with the best-fit
results.
As for the angular correlation function, $w(\theta)$, of detected
SPIRE galaxies, Cooray et al. (2010) reported measurements of
$w(\theta)$ for sources brighter than 30 mJy at all SPIRE
wavelengths and inferred values of $\log(M_{\rm min}/M_\odot)$
ranging from $12.6^{+0.3}_{-0.6}$ at $250\,\mu$m to
$13.5^{+0.3}_{-1.0}$ at $500\,\mu$m. On the other hand, Maddox et
al. (2010) did not detect a significant clustering for their
$250\,\mu$m selected sample with a flux limit of
$33\,\hbox{mJy}\,\hbox{beam}^{-1}$, but detected strong clustering
at $350\,\mu$m and $500\,\mu$m, albeit with relatively large
uncertainties. Our model entails a relationship between the
far-IR/sub-mm luminosity of proto-spheroidal galaxies (that provide
the dominant contribution to $w(\theta)$, see the left-hand panel of
Fig.~\ref{fig:wtheta}) and the associated halo masses (Lapi et al.
2011). For the flux density limit adopted by Cooray et al. (2010),
30 mJy at all SPIRE wavelengths, the model yields $\log(M_{\rm
min}/M_\odot)\simeq 12.3$, 13, and 13.2 at 250, 350, and
$500\,\mu$m, respectively, while for the flux density limits of
Maddox et al. (2010; 33, 36, and 45 mJy) we have $\log(M_{\rm
min}/M_\odot)\simeq 12.3$, 13.1, and 13.4. The corresponding
predictions for $w(\theta)$ are compared with the data in
Fig.~\ref{fig:wtheta}. The agreement of the model with the data is
generally good, although the situation at $250\,\mu$m is unclear
since there is a discrepancy between the Cooray et al. (2010) and
the Maddox et al. (2010) results.
\section{Discussion and conclusions}\label{sect:conclusions}
According to the Granato et al. (2004) model, the steep portion of
sub-mm counts is dominated by massive proto-spheroidal galaxies in
the process of forming most of their stars on a timescale varying
with halo mass (shorter for more massive galaxies), but typically of
$\simeq 0.7\,$Gyr, i.e. with a duty cycle of $\simeq 0.2$ at
$z\simeq 2$, where their redshift distribution peaks. As shown
above, this model allows us to reproduce the power spectrum of CIB
fluctuations over a broad frequency range, from $250\,\mu$m to a few
mm, with only 2 free parameters. The model also yields an effective
volume emissivity at different redshifts consistent with
observational estimates. The derived effective halo mass, $M_{\rm
eff} \simeq 5\times 10^{12}\,M_\odot$, is close to that estimated
for the most efficient star-formers at $z\simeq 2$.
The multipole number at which the 1-halo term starts exceeding the
2-halo contribution to the clustering power spectrum ranges from
$\ell=1450$ at 857 GHz to $\ell=2100$ at 217 GHz. These values are
almost a factor of 2 higher (and, consequently, the corresponding
angular scales are almost a factor of 2 lower) than those found by
Planck Collaboration (2011). Since, at the redshifts where the
contribution to the CIB intensity peaks, the corresponding masses
are well above $M_\ast$, this difference translates into a much
larger abundance of the relevant halos.
Alternative models make quite different predictions for the
clustering properties of sub-mm galaxies. A widespread view is that
these objects are powered by major merger events. Two major theories
have been worked out in this general framework. One view is that
sub-mm galaxies are massive objects, seen during a short-duration,
intense, merger-induced burst of star formation (e.g. Narayanan et
al. 2009). Since massive galaxies are rare at high-$z$ and, because
of the short duration of the burst, only a small fraction of them
are in the sub-mm bright phase at a given time, this scenario has
difficulty in reproducing the observed counts. This difficulty may
be overcome assuming an extremely top-heavy initial stellar mass
function that would allow much less massive (hence far more
abundant) galaxies to reach the required luminosities (e.g. Baugh et
al. 2005; Lacey et al. 2010).
The clustering implied by the latter scenario has been investigated
by Almeida et al. (2011) who found, at $z = 2$, a comoving
correlation length of $r_0 = 5.6\pm 0.9\,h^{-1}\,$Mpc for galaxies
with $850\,\mu$m flux densities brighter than 5 mJy or an effective
bias factor $b_{\rm eff}=2.3$; for galaxies with $S_{450\mu{\rm
m}}>5\,$mJy they found $b_{\rm eff}=2.1$. Our model implies
$\log(M_{\rm min}/M_\odot)\simeq 12.4$ for sources with
$S_{450\mu{\rm m}}>5\,$mJy and $\log(M_{\rm min}/M_\odot)\simeq
13.0$ for sources with $S_{850\mu{\rm m}}>5\,$mJy. The corresponding
values of the clustering radius and of the effective bias factor are
$r_0 \simeq 11.2\,h^{-1}\,$Mpc, $b_{\rm eff}=4.3$ at $850\,\mu$m,
and $r_0 \simeq 7.3\,h^{-1}\,$Mpc, $b_{\rm eff}=3.1$ at $450\,\mu$m.
The study by Kim et al. (2011) confirms that the clustering data
require a higher amplitude of the 2-halo term, i.e. more massive
halos than implied by the major mergers plus top-heavy initial
stellar mass function scenario.
Dav\'e et al. (2010) investigated the clustering properties of
rapidly star-forming galaxies at $z\simeq 2$ in the framework of a
very different scenario based on cosmological hydrodynamic
simulations whereby the star formation is not powered by mergers but
by steady gas accretion and cooling that can fuel the star formation
for several Gyrs. In this scenario typical sub-mm galaxies at $z =
2$ live in massive ($\sim 10^{13}\,M_\odot$) halos and have a duty
cycle $\simeq 50\%$. They are expected to be strongly clustered,
with a clustering radius $r_0\sim 10\,h^{-1}\,$Mpc and a bias factor
of $\sim 6$. These values are well in excess of those following from
our analysis which yields, for the bulk of galaxies at $z\simeq 2$,
$r_0 \simeq 6.9\,h^{-1}\,$Mpc, $b_{\rm eff}\simeq 3$.
These results illustrate the power of accurate measurements of the
CIB power spectrum and of the correlation function of galaxies at
(sub-)millimeter wavelengths to discriminate among competing
evolutionary models for the population of dusty galaxies.
\section*{Acknowledgments}
Thanks are due to G. Lagache for clarifications on the CIB power
spectra derived from {\it Planck} data and to the referee for a
careful reading of the manuscript and useful comments. Our numerical
analysis was performed on the Deepcomp 7000 system of the
Supercomputing Center of Chinese Academy of Sciences. We acknowledge
financial support from ASI (ASI/INAF Agreement I/072/09/0 for the
Planck LFI activity of Phase E2) and MIUR PRIN 2009. MV is supported
by ASI/AAE, PD-INFN 51, PRIN INAF and the FP7 cosmoIGM grants.
|
\section{Introduction}
Substantial progress has been made in
recent years by using lattice methods to solve QCD and, thus, to calculate the spectrum of
baryonic and mesonic excited states from first principles.
In this report, we describe a recent solution to the problem of spin
identification for baryonic excited states. This development has made it possible
to determine patterns in the lattice spectra of $N$ and $\Delta$ excited states
that have the same spins and numbers of states as in
$SU(6)\otimes O(3)$ symmetry.
The lattices and methods used to identify spins are described fully in Ref.~\cite{Edwards:2011} and the reader
should consult that paper for details about the calculations.
The standard method of spin identification relies on patterns of degenerate states across
irreps of the octahedral group in the continuum limit. For
half-integer spins one uses the double-covered octahedral group.
This method has been found to be ambiguous for two simple reasons. First, there is a high degree of degeneracy in lattice spectra and second, the energies of states are subject to uncertainties owing to fluctuations of the gauge configurations. A typical consequence is that approximately degenerate baryon states in $G_1$, $H$ and $G_2$ irreps could indicate a
spin $\frac{7}{2}$ state, but it could equally well indicate an accidental degeneracy of a
spin $\frac{1}{2}$ state and a spin $\frac{5}{2}$ state. This ambiguous situation is not likely to improve
without a new method.
The key developments that have enabled spin identification
and, thus, better insight into
the lattice spectra are i.) the construction of lattice operators
that transform as irreducible
representations (irreps) of the $SU(2)$ rotational symmetry restricted to the lattice, and
ii.) the use of operator overlaps to tell which operators create which states.
The operators are labeled with
a continuum spin quantum number, $J$. They are realized by
incorporating combinations of covariant lattice derivatives that transform as
orbital angular momenta restricted to the lattice. Combinations of the baryon
spin, S, and orbital angular
momenta, L $\leq$ 2, are constructed to provide total spins $J \leq \frac{7}{2}$. A similar
development for integer spins (meson states) was given
in Ref.~\cite{Dudek:2009}.
In order to be suitable for lattice calculations,
operators labeled by continuum spin $J$ must be subduced to lattice operators that transform as irreps of the
octahedral group. Because of the symmetries of a cubic lattice, it is the latter operators that
provide an orthogonal basis for calculations. Subduction is performed using matrices that
provide the change of basis from $SU(2)$ quantum states, labeled as $|J,M\rangle$,
to quantum states that are irreps of the octahedral group in the continuum limit labeled as $|\Lambda, r,[J]\rangle$, where
$\Lambda$ and $r$ denote the irrep and row while $[J]$ denotes the continuum spin. The
matrix elements $\langle J,M | \Lambda,r,[J]\rangle$ provide the subduction matrices.
\section{Approximate rotational symmetry and spin identification}
\begin{figure*}
\hspace{1in} \includegraphics[height=.45\textheight]{matrixplot.pdf}
\caption{Matrix of $H_u$ correlation functions at time slice 5. \label{fig:J_blockdiag} }
\end{figure*}
Even though the lattice spacing used in this work
is not especially small, good evidence is found for an approximate realization of the continuum
rotational symmetry in the lattice spectra. An example of this is given in
Fig.~\ref{fig:J_blockdiag}, which indicates
the relative sizes of matrix elements of a $48\times 48$ matrix of nucleon correlation functions
in the $H_u$ irrep of the double-covered octahedral group. The $u$ subscript of $H_u$ denotes
ungerade, or negative parity. Each pixel indicates the magnitude
of one matrix element according to the scale given in the lower right corner. The 48 operators
used consist of 28 operators that are
subduced from $J=\frac{3}{2}$, whose matrix elements are in the upper block; 16
operators that are subduced from $J=\frac{5}{2}$, whose matrix elements are in the
middle block; and 4 operators that are subduced from $J=\frac{7}{2}$ in the lower block.
Matrix elements that are outside the blocks involve operators subduced from different $J$ values.
They are all quite small. Thus, the matrix is close to
block diagonal in spin, $J$, which is a signal of the approximate realization of rotational
symmetry.
Approximate rotational symmetry also is clearly evident in spectra that are obtained
from matrices of correlation functions. When the matrices are diagonalized, they may
be expressed (at large time t) as a spectral decomposition,
\begin{equation}
C_{ij}(t) = \sum_n \frac{Z_i^{n*} Z_j^{n}}{2m_n} e^{-m_n t},
\label{spectro_decomp}
\end{equation}
where $i$ and $j$ are labels of the operators ${\cal O}_i$ and ${\cal O}_j$ that are involved in the correlation function, the sum is over eigenstates with $m_n$ being the mass of the $n^{th}$ eigenstate and
``overlap factors", $Z^n_i \equiv \langle n | {\cal O}_i^\dag | 0 \rangle$
are matrix elements for the $i^{th}$ operator to create the
$n^{th}$ eigenstate.
The masses $m_n$ determine the spectrum and the overlap factors show which operators create which states.
A detailed examination of the overlap factors shows that lattice states are created
predominantly by operators subduced from a
single continuum spin, $J$. We identify the spin of each lattice state as the
continuum spin of those operators.
\begin{figure*}
\hspace{0.6in}\includegraphics[height=0.58\textwidth,width=0.7\textwidth,bb=20 30 420 396]{ops_test.pdf}
\caption{Extracted Nucleon $H_u$ mass spectrum for various operator bases. Column 1 uses all
48 $H_u$ operators, columns 2, 3 and 4 use restricted bases of 28 $J=\frac{3}{2}$ operators,
16 $J=\frac{5}{2}$ operators and 4 $J=\frac{7}{2}$ operators, respectively. States with spin
identified as $J = \frac{3}{2}$ are shown in red, $J = \frac{5}{2}$ are shown in green and
$J = \frac{7}{2}$ are shown in blue. Results are from the $m_{\pi} $ = 524 MeV ensemble. }
\label{fig:vary_ops}
\end{figure*}
Once spins have been identified, a further test of the approximate rotational invariance
in the lattice spectra becomes possible.
If indeed the couplings are small between states subduced from different continuum spins,
then the omission of such couplings should not much affect the excited state spectra.
That proposition can be tested by calculating energies using all operators,
and comparing them with the energies obtained from the subset of operators subduced
from a single $J$ value.
If approximate rotational invariance is achieved in the spectrum, the
energies should be nearly the same. In Fig.~\ref{fig:vary_ops}, we show such a comparison
for the states in the Nucleon $H_u$ irrep. The left column of Fig.~\ref{fig:vary_ops},
labelled ``all'', shows the lowest 12 energy levels obtained from matrices
of correlation functions using the set of all 48 $H_u$ operators, with spins identified using the
overlaps, as described above. The second column shows the lowest 6 levels resulting from the variational method when the operators are subduced from continuum
spin $J=\frac{3}{2}$.
Similarly, the third column shows the lowest 4 levels obtained from the operators subduced from continuum spin $J=\frac{5}{2}$, and the last column shows the lowest two levels from operators subduced from $J=\frac{7}{2}$.
The results are striking.
We see that the masses of the levels in each of the restricted bases agree quite well with the results
found in the full basis.
The agreement is quite remarkable because one expects that
operators in the $H_u$ irrep that can couple with the ground state (the lowest
$J=\frac{3}{2}$ state), will show a rapid decay of correlators down to the ground state as
a function of time, $t$. However, the
higher-energy spin $\frac{5}{2}$ and $\frac{7}{2}$ states do not show such a
decay; we obtain good fits to correlators at large $t$, where a single mass dominates the decay.
These results provide a rather striking demonstration for the lack of significant rotational symmetry
breaking in the spectrum.
\section{Lattice $N^*$ spectrum with spins identified}
The robust identification of spins allows the
spectrum to be displayed as in Fig.~\ref{fig:J_spectrum_bands}, with masses and uncertainties
indicated by boxes in columns labeled by the spin and parity values of the states, $J^P$.
The excited nucleon spectrum of Fig.~\ref{fig:J_spectrum_bands} contains many more states than
the experimental spectrum. All the lattice states, even multiparticle states, are discrete levels (within
statistical uncertainties) because of the use of a finite box.
Moreover, the use of three-quark operators, as in this work, does not give significant multiparticle
contributions, although they should be present in principle. Thus, the expected washing out of highly excited states
when they are embedded in a continuum of multiparticle states, does not show up with a small
lattice volume. The lattice spectrum is expected to become more realistic
with multiparticle operators included in the analyses and volume larger than the present value of $(2 \rm{fm})^3$.
In Fig.~\ref{fig:J_spectrum_bands}, bands of states are
indicated by the shaded regions. The lowest band has negative parity states, the next higher band
has positive parity states, and as the mass increases, the bands alternate from one
parity to the other in a staircase fashion. It has been suggested that a restoration of chiral symmetry might
occur at high excitation energy in QCD. ~\cite{Glozman:1999}
That would show up as close to identical masses of higher states in both
positive- and negative-parity spectra.
The lattice spectra at $m_{\pi} = $396 MeV do not provide support for such a parity doubling.
\section{$N^*$ states with the quantum numbers of $SU(6)\times O(3)$ states }
Focusing on the lowest band of $N^*$ states, Fig.~\ref{fig:J_spectrum_band1}
shows that the five states have spins based on the combinations of $S=\frac{1}{2}$ and
$S= \frac{3}{2}$ with orbital angular momentum and parity $L=1^-$. The number of states is tabulated
inside the shaded region for each $J$ value and below that we tabulate the
construction of spins as $S \oplus L \rightarrow J$.
These 5 states belong to the $[70, 1^-]$ multiplet of $SU(6)\times O(3)$. Although
the nucleon operators used are based on four-component Dirac spinors for each quark, the
lowest-energy
states are created predominantly by operators with only upper components of the Dirac spinors, i.e., Pauli spinors.
The next band at higher mass and positive parity contains 13 $N^*$ states in the shaded region of Fig.~\ref{fig:J_spectrum_band2}, with the number of states listed for each $J$ value. Below that
is a tabulation showing the $S \oplus L \rightarrow J$ constructions of the 13 states.
The operators used allow the construction to be determined in the same fashion as the spin,
i.e., by the overlaps.
As can be seen in the tabulation, the 13 states correspond to the positive parity states that can be made by combining S = $\frac{1}{2}$
and $\frac{3}{2}$ with $L = 0^+, 1^+$ and $2^+$. Nonzero L values are incorporated by covariant derivatives in the
operators. If the derivatives were omitted, only one $J= \frac{1}{2}^+$ state would appear in
the shaded region rather than
the 4 states that are obtained with derivatives. That is interesting because previous lattice analyses that omit
derivative operators generally have found one or two close-together excited $\frac{1}{2}^+$ states rather than 4
close-together states, as are obtained here. Our results suggest that the Roper resonance
could have a more complex structure that would not be seen without the use of derivative operators.
It is an interesting question whether this pattern will still be evident when multiparticle
operators are included in the analysis.
It has been suggested that a quark-diquark model might account for the $N^*$ spectrum.
The lattice results show that the $[20, 1^+]$ multiplet is realized, i.e., it gives the 5
$L = 1^+$ constructions seen in the tabulations in Fig.~\ref{fig:J_spectrum_band2}. Those states
are not consistent with a quark-diquark spectrum.
\section{Summary}
The identification of spins of lattice states is significant because the usual method gives ambiguous
results, which is a consequence of the high degree of degeneracy in the spectra and uncertainties
in the determinations of lattice energies.
The use of operators subduced from continuum spins
allows robust spin identification.
That, in turn, allows
the interpretation of low-lying states in the lattice spectra in terms of the quantum numbers
of $SU(6)\otimes O(3)$ symmetry. The clarity of the spectrum with spins identified
provides contraindications with regard to three oft-discussed issues: parity doubling, quark-diquark models and
a simple structure of the $N^* \frac{1}{2}^+$ (Roper) resonance.
Support from U.S. Department of Energy contract DE-FG02-93ER-40762 is acknowledged.
Computations were performed using {\tt Chroma}~\cite{Edwards:2004sx}
and {\tt QUDA}~\cite{Clark:2009wm,Babich:2010mu} at Jefferson Laboratory
under the USQCD Initiative and LQCD ARRA project.
\begin{figure*}
\vspace{-.3in}\hspace{0.5in}\includegraphics[width=0.8\textwidth]{nucleon_Jg_Ju_bands.pdf}
\caption{Lattice $N^*$ energy spectrum in columns of good $J^P$ at $m_{\pi}$ = 396 MeV with spins identified. Shaded regions
show bands of states with alternating parity. \label{fig:J_spectrum_bands}}
\end{figure*}
\begin{figure*}
\vspace{-.3in} \hspace{0.7in}\includegraphics[width=0.75\textwidth]{nucleon_Jg_Ju_SU6O3-1.pdf}
\caption{Five $N^*$ states corresponding to the $[70, 1^-]$ multiplet of $SU(6)\times O(3)$ are
shown in the shaded area.
The $J$ values and parity are the same as can be formed by $S+ L \rightarrow J$,
as tabulated in red, providing the same numbers of states of each spin
as in the lattice spectrum
\label{fig:J_spectrum_band1}}
\end{figure*}
\begin{figure*}
\hspace{.7in}\includegraphics[width=0.75\textwidth]{nucleon_Jg_Ju_SU6O3-2.pdf}
\caption{Thirteen $N^*$ states corresponding to the $[56,0^+]$, $[56,2^+]$ $[70,0^+]$ $[70,0^+]$ and $[20,1^+]$
multiplets of $SU(6)\times O(3)$ are shown in the shaded area.
The $J$ values and parity are the same as can be formed
by $S + L \rightarrow J$, as tabulated in red, when orbital angular momenta and parity
$L= 0^+, \ 1^+,\ $or $2^+$ are added to the possible quark spins, providing
the same numbers of states of each spin
as in the lattice spectrum. \label{fig:J_spectrum_band2}}
\end{figure*}
|
\section{Introduction}
The tau lepton, with a mass of $1776.82\pm0.16$~MeV~\cite{PDG}, is the only lepton heavy
enough to decay both leptonically and hadronically. It decays approximately 65\% of the time to
one or more hadrons and 35\% of the time leptonically. The reconstruction and identification of
tau leptons are important in many searches for new phenomena~\cite{CSC}.
Standard Model processes such as $W\rightarrow\tau\nu$, $Z\rightarrow\tau\tau$ boson production can
also result in signatures with tau leptons, which can be used to
measure key quantities such as the tau lepton identification efficiency.
A challenge in identifying hadronic tau decays ($\tau_\mathrm{had}$)
is to distinguish them from hadronic jets.
However, $\tau_\mathrm{had}$ leptons possess
certain properties that can be used to differentiate them from jets.
They usually decay into one (1-prong) or three (3-prong) charged particles
and their decay products are well collimated.
The tau lepton proper lifetime is 87~$\mu$m, leading to decay
vertices that can be resolved in the silicon tracker from the primary interaction vertex.
\section{Tau Reconstruction and Identification}
In ATLAS~\cite{DetPap}, hadronically decaying tau candidates~\cite{taureco} are seeded by anti-$k_\mathrm{t}$
jets~\cite{antikt} with distance parameter $R=0.4$ satisfying $p_\mathrm{T}>10$ GeV and $|\eta|<2.5$ built
from three dimensional clusters of calorimeter cells~\cite{topocluster}.
Tracks reconstructed with $p_T > 1$ GeV and within $\Delta R < 0.2$ of the jet seed are associated to the tau
candidate if they satisfy cuts on the impact parameter and minimum silicon hit criteria.
The energy scale of tau candidates is determined using a two-step process. In the first step,
local hadron calibration~\cite{LCcalib} is applied to clusters
within a radius of $\Delta R < 0.2$ of the barycentre.
The resultant energy from the sum of cluster four-vectors is used for the second step of the calibration.
In this step, an additional correction on this energy is applied, based on
Monte Carlo (MC) studies of processes involving hadronic tau decays, to obtain the fully calibrated
tau candidate energy. Uncertainties on the energy scale are determined by comparing
the calibrated energy in different MC simulation samples, with realistic variations
of conditions such as the hadronic shower model and dead material modelling~\cite{taureco}.
Discrimination against background candidates from jets and electrons is provided in a separate identification step.
Identification variables are reconstructed from tau candidates, based on tracker and calorimeter
information~\cite{taureco}. Examples of such identification variables include: the core energy fraction
($f_\mathrm{core}$), the ratio of energies at the electromagnetic scale deposited within
$\Delta R < 0.1$ and $\Delta R < 0.4$ of the tau candidate; and the transverse flight path
significance ($S_\mathrm{T}^\mathrm{flight}$),
the transverse decay length significance of the reconstructed vertex of multi-prong candidates.
Distributions of $f_\mathrm{core}$ for 1-track candidates and
$S_\mathrm{T}^\mathrm{flight}$ for 3-track candidates are shown in Figure~\ref{fig:IDvars}, for
both signal (in simulated $W\rightarrow\tau\nu$ and $Z\rightarrow\tau\tau$ events) and
jet background (from a dijet selection) candidates.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{fig/f_core.eps}
\includegraphics[width=0.45\textwidth]{fig/s_flight.eps}
\caption{
Two discriminating variables used for the tau identification: $f_\mathrm{core}$ and $S_\mathrm{T}^\mathrm{flight}$~\cite{taureco}.
The filled histograms are from MC simulation, the points are data with a dijet selection.
\label{fig:IDvars}}
\end{figure}
The identification variables are combined into discriminants to suppress background candidates from
jets and electrons. ATLAS has developed three such discriminants: a cut-based discriminant,
a projective likelihood, and a boosted decision tree~\cite{taureco}. The performance of these
discriminants can be evaluated by plotting the rejection (inverse efficiency) for jets in a dijet
selection against the efficiency for signal tau candidates, for both 1-track and 3-track candidates.
This is shown in Figure~\ref{fig:Performance}.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{fig/perf_1p.eps}
\includegraphics[width=0.45\textwidth]{fig/perf_3p.eps}
\caption{
Inverse background efficiency in dijet data as a function of signal efficiency
in $W\rightarrow\tau\nu$ and $Z\rightarrow\tau\tau$ MC events for all
discriminants on 1-track and 3-track candidates~\cite{TauWiki}.
\label{fig:Performance}}
\end{figure}
\section{Tau Identification Efficiency Measurements}
The identification efficiency for hadronic tau decays is measured with $W\rightarrow\tau\nu$ events~\cite{WtaunuEff}.
In the tag \& probe method, events are selected with
$S_{E_\mathrm{T}^\mathrm{miss}} = E_\mathrm{T}^\mathrm{miss} / (0.5 \mbox{GeV}^{1/2}\sqrt{\Sigma E_\mathrm{T}}) > 6$,
where $E_\mathrm{T}^\mathrm{miss}$ is the missing transverse energy and $\Sigma E_\mathrm{T}$ is the scalar sum of
cluster transverse energy.
A tau candidate that is well separated from the direction of the $E_\mathrm{T}^\mathrm{miss}$ is required,
and the tau track multiplicity is fitted before and after applying tau identification to extract the fraction of
$W\rightarrow\tau\nu$ events in the sample. The track multiplicity templates for true $\tau_\mathrm{had}$ leptons
are taken from $W\rightarrow\tau\nu$ MC events, while for jet candidates, a template from a jet-enriched
control region $2 < S_{E_\mathrm{T}^\mathrm{miss}} < 4.5$ is used.
The tau track multiplicity before and after tau identification is shown in Figure~\ref{fig:WtaunuTrack}.
\begin{figure}
\centering
\includegraphics[width=0.40\textwidth]{fig/Wtaunu_ntrack_before.eps}
\includegraphics[width=0.40\textwidth]{fig/Wtaunu_ntrack_after.eps}
\caption{
Track multiplicity before and after tau identification~\cite{WtaunuEff}. The hatching represents the systematic uncertainty.
The normalisation of the different processes is determined through a fit to the track multiplicity spectrum.
\label{fig:WtaunuTrack}}
\end{figure}
This measurement is cross-checked with a second method that compares the observed $W\rightarrow\tau\nu$
event yield with the predicted $W\rightarrow\tau\nu$ event yield based on the measured
$W\rightarrow e\nu$ and $W\rightarrow \mu\nu$ cross-sections.
Both methods measure a tau identification efficiency in $W\rightarrow\tau\nu$ events that are
consistent with the predicted efficiency from MC simulation~\cite{WtaunuEff}.
\newpage
\section{Summary}
ATLAS has developed a well-performing reconstruction algorithm to identify hadronic tau decays, enabling
various measurements and searches of physics processes with tau leptons in the final state. MC predictions of the
identification efficiency are shown to be consistent with measurements of the efficiency in
$W\rightarrow\tau\nu$ data events.
|
\section{Introduction}
Graphene-based systems have recently attracted much interest from both experimental and theoretical aspects.
Its chemical inertness, hydrophobic behavior, large electron mobility, scalable production and intrinsically low spin-orbit coupling makes it a very promising candidate for sensors\cite{schedin}, electronics \cite{Neto}, spintronics \cite{{tombros},{Candini2011}} and nanomechanics \cite{{Bunch},{Bachtold}}.
Experimentally, the graphene capability to detect small charge transfer effects by grafting molecules\cite{{schedin},{haddon}}, but also to induce superconductivity\cite{Kessler,Heersche}, as well as Kondo effect\cite{Krasheninnikov} by deposing metals have been investigated.
More recently, surface-enhanced Raman signal has been realized on pyrene-based molecules grafted on graphene \cite{{Dress}, {Bendiab}}, allowing a detection down to the limit of few isolated molecules.
Detection of single molecule magnet\cite{{Bendiab},{matias},{WWreview}} by using graphene or carbon nanotubes could also be a way to probe magnetic properties at the single molecule level.
Theoretically, \textit{ab initio} calculations on magnetic properties of graphene based materials have been performed\cite{Xiao,Leenaerts}.
In particular, \textit{ab initio} studies of adatoms on graphene \cite {Johl,Mao,Nieminen2003} magnetic Co dimers\cite{Xiao} and even small magnetic molecules such as O$_2$, NO or NO$_2$\cite{Leenaerts} have been reported.
The interaction between the graphene monolayer and a magnetic molecule is an important point in order to have insights on local effects such as induced magnetization and charge transfer, which are of interest for spintronic and nano-electronic applications.
In the present work, we chose the iron tetraphtalic acid molecule - $FeTPA_{4}$ \cite{Gambardella} - deposited on monolayer graphene as a case study of this interaction.
Indeed, Gambardella et al. \cite{Gambardella} have succeeded in the manipulation of the magnetic anisotropy of a supramolecular assembly of $FeTPA_{4}$ self assembled on a Cu surface.
However, to speed up our DFT calculations, we modeled $FeTPA_{4}$ by explicitly studying its magnetic core $Fe(OH)_{4}$, as we will discuss in the following, where this choice will be justified on the basis of structural, electronic, and magnetic properties.
\section{Computational details}
First principles DFT calculations were performed within the plane wave approximation (PW) and pseudopotential scheme, as implemented in the quantum ESPRESSO code \cite{QE, PW}.
We adopt a Perdew-Burke-Ernzerhof (PBE) \cite{Perdew} gradient corrected functional for the exchange and correlation potential, and ultrasoft pseudopotential technique is used to describe C, H, O and Fe atoms \cite{LS}.
Electronic wave functions are written in terms of plane waves, with an energy up to 30 Ry, which is sufficient to ensure convergence of structural, electronic and magnetic properties.
The electronic occupation is computed using Fermi-Dirac distribution, with a smearing parameter of 136 meV, which corresponds to an electronic temperature of 1578 K.
The Brillouin zone integration is performed with a uniform \textbf{k} points grid of (15x15x1).
The convergence of the electronic and spin properties with the electronic temperature and k-points grid has been tested.
Our choice of 1578 K with (15x15x1) k-points grid is within the range of typical electronic temperatures used to describe graphitic systems, and it is motivated by the fact that such values ensure converged results with reasonable computational effort.
Structures are relaxed until forces are below 0.02 eV/\r{A}.
We used a supercell composed by (6x6) repetition of a graphene unitary cell in order to ensure a negligible interaction between
periodic images of the adsorbed molecule.
Spin polarized calculations are also performed, and we consider polarization along the z axis (\textit{i.e.} perpendicular to graphene plane). \\
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=9cm]{Figure1.pdf}
\caption{(Top and middle panels) Structural description of: a) $Fe(TPA)_4$ isolated molecule, b)
$Fe(OH)_4$ isolated molecule, c) hybrid system composed of graphene and $Fe(OH)_4$. Iron and oxygen atoms are indicated in red,
hydrogen atoms in blue, and carbon atoms in yellow. The high symmetry adsorption sites are indicated with respect
to the hexagonal cell of graphene, (d). Bottom panel: charge density map within xy plan for e)
$Fe(TPA)_4$ and f) $Fe(OH)_4$.}
\label{fig:Figure1}
\end{center}
\end{figure}
\subsection{Modeling $FeTPA_{4}$ through a smaller molecule}
The $FeTPA_{4}$ molecule is difficult to simulate due to its large number of atoms (33 atoms) and, especially, of the size of the graphene supercell that should contain it.
As mentioned above, however, the $Fe(OH)_4$ molecule mimics well the $Fe(TPA)_4$ molecule.
This smaller, case-study system contains the same $FeO_4$ core as the bigger one, but the four large pyrenic arms are simply replaced by hydrogen atoms.
As shown in the top panels of Fig. \ref{fig:Figure1}, the two molecules have a similar planar structure.
The charge distribution around the iron core are computed projecting all the electronic wave functions $\varphi_{i,\sigma}$ on the atomic orbitals centered in each given atom through the equation:
\begin{equation}
n^{\beta} = \sum _{i, \ell, m,\sigma=\uparrow,\downarrow} \left|\left\langle \phi_{\ell,m}^{\beta} | \varphi_{i,\sigma} \right\rangle \right|^{2},
\label{eq:N_projection}
\end{equation}
where $\phi_{n \ell}^{\beta}$ is the atomic-like orbital for atom $\beta$, labelled with its energy and angular momentum quantum number $m,l$.
\begin{table}[htbp]
\begin{center}
\begin{tabular*}{8cm}{@{\extracolsep{\fill}}clcccc}
\hline
\hline
Molecule & Atom & \multicolumn{2}{c}{Lowdin charge} & Polarization \\
\hline
\hline
$Fe(OH)_{4}$ && & &\\
&Fe &15.628 & &3.196 \\
& &\textit{s} &2.503 &\\
& &\textit{p} &6.722 &\\
& &\textit{d} &6.402 &\\
&O &6.450 & &0.186\\
& &\textit{s} &1.686 &\\
& &\textit{p} &4.764 &\\
\hline
$Fe(TPA)_{4}$ && & &\\
&Fe &15.611 & &3.102\\
& &\textit{s} &2.473 &\\
& &\textit{p} &6.741 &\\
& &\textit{d} &6.395 &\\
&O* &6.246 & &0.122\\
& &\textit{s} &1.623 &\\
& &\textit{p} &4.623 &\\
\hline
\hline
\end{tabular*}
\caption{Calculated Lowdin charges (see eq.\ref{eq:N_projection}) and their
decomposition onto atomic-like orbitals, for
Fe and O atoms in $Fe(OH)_4$ and $Fe(TPA)_4$ molecules. O* stands for first neighbors of Fe atom.
Polarization is computed along the z direction for Fe and O atoms as the difference between spin up and down charges.}
\label{charge_proj}
\end{center}
\end{table}
We observe in Table \ref{charge_proj} that the charge associated to the core iron atom (Lowdin charge) remains very similar in both cases.
The only slight charge difference between the two cases concerns the O atoms in $Fe(OH)_4$ having around 0.2 electrons more
than in $Fe(TPA)_4$.
We also computed the spin polarization along the $z$ direction for Fe and O atoms, as the difference between the spin up and down charges.
As shown in Table \ref{charge_proj}, the spin polarization of Fe atoms is very similar, being 3.20 and 3.10 $\mu_B$ in $Fe(OH)_4$ and $Fe(TPA)_4$ molecules respectively.
In both cases this polarization comes from unpaired electron in 3d iron orbitals.
The polarization of the O atoms, which comes from 2p orbitals, is slightly higher in the $Fe(OH)_4$ molecule.
The total polarization of the core is 3.94 and 3.59 $\mu_B$ in $Fe(OH)_4$ and $Fe(TPA)_4$, respectively, which corresponds to a variation of around 10$\%$.
We can thus safely consider these two molecules equivalent from the spin magnetic point of view in the void; given the relative inertness of graphene sheets with respect to full metallic surfaces, we can reasonably extrapolate a very similar behavior when grafted on a graphene sheet.
\subsection{Calculation details of the hybrid system}
We use, to study the hybrid system composed by the $Fe(OH)_4$ molecule adsorbed on graphene, an hexagonal supercell composed of 36 unit cells (Fig.\ref{fig:Figure1}c) with a lattice parameter in the xy plane of a$_0$ = 14.76 \r{A}, and of 11.10 \r{A} in z direction.
Due to Brillouin zone refolding, the Dirac (K) point of graphene refolds on the $\Gamma$ point.
The three high symmetry adsorption sites of $Fe(OH)_4$ on graphene are shown in Fig.\ref{fig:Figure1}d.
We find that the most energetically favorable adsorption site is the top site, and only this one will then be considered in the following.
\section{Results and discussion}
\subsection{Structural properties}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Figure2.pdf}
\caption{(a) Binding energy of the hybrid system as a function of the distance along z between the graphene surface and the iron atom
of the Fe(OH)$_{4}$ molecule. (b) Displacements ($\Delta z^{C\#} $) of graphene C atoms positions along z with respect to the
isolated case, as a function of distance, within the xy plane, to the adsorption (top site) C$_{top}$ atom. The
horizontal arrow represents the planar extension of $Fe(OH)_4$ molecule.
Two fixed values of z$_{Fe-G}$ are considered : the equilibrium distance
4.1 \r{A} (filled circles) and the characteristic distance 3.1 \r{A} (empty
triangles)}
\label{fig:Figure2}
\end{center}
\end{figure}
We performed the structural relaxation of the hybrid system by setting a fixed distance z$_{Fe-G}$ between the Fe atom in the $FeOH_{4}$ molecule and the graphene plane.
According to our calculations, top adsorption site is found to be the most stable from an energetic point of view and we defined z$_{Fe-G}$ as the distance along z between the Fe atom and the first C atom below, which we indicate as C$_{top}$.
We calculated the binding energy as a function of this distance (see Fig. \ref{fig:Figure2}), as $E_{binding} = E_{graphene + FeOH_{4}} - E_{graphene} - E_{FeOH_{4}}$.
The obtained equilibrium distance for the system is z$_{Fe-G}^{eq}$=4.1 \r{A}, and the corresponding binding energy is $E_{binding} = -0.36$ eV.
However, it is well known that \textit{ab initio} DFT calculations do not take into account Van der Waals interactions, which are of course important in
weakly interacting systems.
As a result, the distance we obtain is likely to be overestimated.
A properly \textit{ab initio} correction should explicitly take into account these contributions in the DFT functional.
However, this procedure is rather cumbersome and computationally costly, and a common approximation consists in carrying out the calculations at the experimental equilibrium distance.
The most flagrant case is graphite, a 3D crystal formed by layers of graphene, whose experimental interplane equilibrium distance is about 3.35 \r{A}, while the DFT equilibrium value is more than 5 \r{A}.
In an analogous way, in the following we always compare results obtained for the DFT equilibrium distance z$_{Fe-G}^{eq}$=4.1 \r{A} with results obtained at a shorter distance, z$_{Fe-G}$= 3.1 \r{A} which is comparable to the characteristic interaction distance observed in sp$^{2}$ carbon materials \cite{cardona,dresselhaus}.\\
In both cases, the flatness of the graphene plane is slightly perturbed by the presence of the molecule.
While C-C bond distances are practically unaffected ($\leq$ 0.1 $\%$), the z coordinate of C atoms in the relaxed hybrid system changes by as much as 1 $\%$, as shown in Fig. \ref{fig:Figure1}.
This surface modulation extends up to a distance of around 5.5 \r{A} whether z$_{Fe-G}$ equals 4.1 or 3.1 \r{A}.
\subsection{Electronic properties}
The charge transfer between graphene layer and the molecule can be quantified by computing the difference between the total charge of the graphene sheet in the hybrid system and of isolated graphene.
We obtain a charge difference of -0.63 electrons, revealing an induced hole doping on graphene.
This charge transfer is equal to a surface charge density of about -3.4 10$^{13}$ cm$^{-2}$, which is coherent with common predictions and experiments involving graphene \cite{Novoselov2, Novoselov3, Ferrari}.\\
We also evaluate the local electronic charge transfer for each C atoms in the graphene plane and extract contributions from the different orbitals by projecting the electronic eigenstates on the atomic wavefunctions as in Eq. \ref{eq:N_projection}.
At a molecule-graphene distance of 4.1 \r{A} we observe that the charge transfer from the graphene sheet originates mostly (98\%) from the 2p orbitals of C atoms.
A maximum charge transfer is observed for 2p states of C atoms at 2.5 \r{A} from the Fe center, while at distances greater than 6 \r{A} the spatial charge transfer remains constant.
For $z_{Fe-G} = 3.1 $\r{A}, a similar behavior is observed.
From this result we determine an interaction length of about 6 \r{A}, which is in agreement with the one estimated by the analysis of the surface corrugation (see Fig.\ref{fig:Figure1}).
Moreover, the comparison between the electronic band structure of isolated graphene and of the hybrid system (not reported) indicates that, despite a shift of the Fermi level, the perturbation induced by the molecule is negligible, as also shown by the non-dispersive character of the molecular states. \\
This analysis of the charge transfer and of the electronic band structure of the hybrid system suggests that the quasi-metallic character of graphene is not perturbed by its interaction with the grafted magnetic molecule.
This result is promising in view of the use of graphene as a sensitive detector.
\subsection{Magnetic properties}
As mentioned above, we also performed spin polarized calculations on isolated molecule and on the hybrid system.
The obtained magnetic moment for the isolated molecule $Fe(OH)_{4}$ is about 3.94 $\mu_B$, whereas in the hybrid system it is about 3.34 $\mu_B$.
The magnetization difference between the two systems is about 0.6 $\mu_B$ indicating that the interaction with the graphene substrate generates 0.6 ``paired'' electrons in this hybrid system which is coherent with the estimated charge transfer of 0.63 electrons found in the previous section.
We then analyze this variation of the molecular magnetization within the hybrid system spin states occupations.
Thus, we calculated the projected density of states (PDOS) for spin up and spin down states within the hybrid system.
More precisely, we focused on the projection into 2p$_z$ states of carbon atoms, which are the most sensitive to their environment.\\
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Figure3.pdf}
\caption{(a) Difference between PDOS (see text) for spin up and down states within the hybrid system.
Projections are performed onto carbon 2$p_z$ atomic orbitals. Carbon atoms
are labelled by their neighbouring rank, in the xy plane, with respect to C$_{top}$.
The sketch represents the C atoms of the graphene. $Fe(OH)_4$ is located at 3.1 \r{A} vertically above the black C$_{top}$ atom and the different colors indicates the different neighboring rank.
$PDOS_{2p_z}(\uparrow) - PDOS_{2p_z}(\downarrow)$ is averaged for every
$C^{\#}$ atom at the same distance from C$_{top}$. (b) Spatial distribution
of the magnetization $\Delta n(\vec{r}) = n(\vec{r})_\uparrow -
n(\vec{r})_\downarrow$ within hybrid system for two distances z$_{Fe-G}$ =
4.1 \r{A} (top) and z$_{Fe-G}$ = 3.1 \r{A} (bottom).} \label{fig:Figure3}
\end{center}
\end{figure}
In Fig. \ref{fig:Figure3}a we report the difference between the PDOS for spin up and down states for different C atoms, labelled with their distance in the xy plane with respect to Fe atom in the $Fe(OH)_{4}$ molecule.
This figure shows an energy range [-3 ; -1.5] eV where the PDOS for spin up and spin down differs and, in particular, for carbon atoms located up to the 5$^{th}$ neighbour (\textit{ie.} 5 \r{A} from C$_{top}$ atom).
This spin-interaction distance is consistent with the previously discussed structural and electronic interaction lengths of the hybrid system.\\
To go further, we analyze the mixing between atomic orbitals of $Fe(OH)_4$ molecule and the ones of graphene in the hybrid system.
In order to quantify the hybridization level of the occupied states, we define the quantity $c_{g/M}$:
\begin{equation}
c_{g/M} = \sum_{AO} c^{AO}_{g,M}
\label{eq:coeff}
\end{equation}
where $C^{AO}_{g,M}$ represents the projection of an electronic states into atomic orbitals (AO) belonging to graphene atoms (g) or to the molecule atoms (M).
These states in the range [-3.2;-2.0] eV, with respect to the Fermi level, correspond to occupied electronic states where $c_{g}/c_{M}\in[10\% ; 90\%]$.
Within that interval, both systems (\textit{i.e.} graphene and molecule) have a significant contribution to the global electronic wavefunctions which suggest a remarkable hybridization.
This is consistent with findings shown in Fig. \ref{fig:Figure3}a, where we observe strongly polarized states in a region between [-3.0 ; -1.5] eV and also with the existence of additional mixed states .
In particular, at \textbf{k}=0 and at -2.4 eV there is an electronic state which is particularly important since it represents the overlap of Fe 3d$_{z^2}$ (13 $\%$) and top C atom 2p$_{z}$ orbital (8 $\%$).
Moreover, the contribution of this atomic orbital from the closest carbon atom is not present in any other mixed wavefunction.
This indicates, for instance, the negligible interaction between the C atom and the oxygen atoms. All other Fe states are only present in pure molecular electronic states and not in mixed states.
Conversely, we do not observe strongly hybridized spin down states.
The hybridized states within [-3.2;-2.0] eV show mainly overlap between oxygen 2p$_{x,y}$ states and 2p$_{z}$ states of carbon atoms located at distances $\geq 5.5 \r{A}$ from Fe atom.
On the basis of our findings, we expect a spin up charge excess around the carbon atoms closer to Fe atom and an excess of spin down charges within a crown of carbon atoms around 5.5 \r{A} from Fe atom.
This local magnetization within graphene plane is shown in Fig. \ref{fig:Figure3}b.
The induced magnetization in the graphene plane is suggested by concentric crowns of spin up/down electron in 2p states of carbon atoms.
\section{Conclusions}
We report a case study of an iron-based magnetic molecule grafted on a graphene sheet and, through \textit{ab initio} DFT calculations, we predict an induced magnetization effect to the graphene substrate.
We also describe a typical interaction length of 5-6 \r{A} within the graphene plane with respect to structural deformations, localized charge transfer and induced magnetization.
The latter originates from a significant coupling between 2p$_z$ and Fe 3d$_{z^2}$ orbital for the closest carbon atom.
On the other hand, we observe that the presence of an iron-based magnetic molecule in the close vicinity of a graphene sheet does not strongly perturb the electronic properties of the latter, thus providing a promising perspective to the design of non-destructive sensor devices.
We acknowledge useful discussions with X. Blase, V. Bouchiat, A. Candini, M. Urdampiletta and W. Wernsdorfer.
Fig. 1 and 3 have been performed with XCrysden package \cite{xcry}.
Calculations were done at IDRIS (Orsay, France), Project No. CP9-91387.
|
\section*{Version fran\c{c}aise abr\'eg\'ee}
Le programme de Kac en th\'eorie cin\'etique consiste \`a comprendre
comment d\'eduire l'\'equation de Boltzmann spatialement homog\`ene
\`a partir d'un processus stochastique de saut sur l'espace des
vitesses \`a grand nombre de particules. Le but de ce programme est de
comprendre la notion de \leavevmode\raise.3ex\hbox{$\scriptscriptstyle\langle\!\langle$~} chaos mol\'eculaire\leavevmode\raise.3ex\hbox{~$\!\scriptscriptstyle\,\rangle\!\rangle$}{} dans un cadre
plus simple que celui de la dynamique compl\`ete des particules, ainsi
que de donner une justification microscopique au th\'eor\`eme $H$
(croissance de l'entropie) et au processus de retour vers
l'\'equilibre. Nous renvoyons \`a la version compl\`ete pour
l'introduction de l'\'equation de Boltzmann \eqref{el}, des processus
de saut consid\'er\'es, ainsi que pour les d\'efinitions de la notion
de chaos et de la distance de Wasserstein $W_1$. \medskip
\begin{theoreme}[R\'esum\'e des r\'esultats principaux]\label{theo:main:fr}
On consid\`ere $d \ge 2$ et une distribution initiale $f_0 \in
P(\mathbb{R}^d) \cap L^\infty$ \`a support compact ou poss\'edant
suffisamment de moments polyn\^omiaux born\'es, et que l'on suppose
centr\'ee sans perte de g\'en\'eralit\'e. Soit $f_t$ la solution
correspondante de l'\'equation de Boltzmann \eqref{el} pour les
sph\`eres dures ou les mol\'ecules maxwelliennes (sans troncature
angulaire), et soit $f^N_t$ la solution du processus de saut \`a $N$
particules correspondant, avec pour donn\'ee initiale $f_0 ^N$: soit
(a) la tensorisation de $f_0 ^{\otimes N}$ de $f_0$, ou (b) la
tensoris\'ee $f^{\otimes N} _0$ conditionn\'ee \`a la sph\`ere
$\mathcal S^N$ (d\'efinie par \eqref{sphere-bol}).
\begin{enumerate}
\item {\bf Propagation de chaos quantifi\'ee et uniforme en temps}:
On consid\`ere le cas (a) o\`u $f_0 ^N = f_0 ^{\otimes N}$. Alors
\begin{equation*
\forall \, N \ge 1, \ \forall \, 1 \le \ell \le N, \quad
\sup_{t \ge 0} {W_1 \left( \Pi_{\ell} f^N
_t, \left( f_t ^{\otimes \ell} \right) \right) \over \ell} \le \alpha(N)
\end{equation*}
avec $\alpha(N) \to 0$ lorsque $N \to \infty$, et o\`u $\Pi_\ell g^N$
d\'esigne la $\ell$-marginale d'une probabilit\'e $g$
sur $(\mathbb R^d)^N$.
\item {\bf Propagation du chaos entropique}: On consid\`ere le cas (b)
o\`u $f_0 ^N$ est conditionn\'ee \`a $\mathcal S^N$. Alors
la solution est entropie-chaotique~:
$$
\forall \, t \ge 0, \quad \frac1N \, H\left( f^N _t | \gamma^N \right)
\to H\left(f_t | \gamma \right), \quad N \to +\infty
$$
(voir \eqref{rel-entropie} pour les d\'efinitions des fonctionnelles
$H$) avec $\gamma$ la probabilit\'e gaussienne centr\'ee d'\'energie
$\mathcal E$ \'egale \`a l'\'energie de $f_0$ et $\gamma^N$ la mesure
de probabilit\'e uniforme sur $\mathcal S^N$. Cela fournit une
d\'erivation microscopique du th\'eor\`eme $H$ dans ce contexte.
\item {\bf Taux de relaxation ind\'ependants du nombre de particules}:
On consid\`ere le cas (b) o\`u $f_0 ^N$ est conditionn\'ee \`a $\mathcal
S^N$. Alors
\begin{equation*
\forall \, N \ge 1, \ \forall \, 1 \le \ell \le N, \ \forall \, t
\ge 0, \quad
{W_1 \left( \Pi_{\ell} f^N
_t, \Pi_\ell \left( \gamma^N \right) \right) \over \ell} \le
\beta(t) \qquad \mbox{avec} \ \beta(t) \to 0, \ t \to 0.
\end{equation*}
Dans le cas des mol\'ecules maxwelliennes, et si la donn\'ee initiale
poss\`ede une information de Fisher finie (voir \eqref{fisher}), on
prouve \'egalement
$
\forall \, N \ge 1, \ 0 \le \frac1N \, H\left( f^N _t | \gamma^N \right)
\le \beta(t) \quad \mbox{avec} \ \beta(t) \to 0, \ t \to 0.$
\end{enumerate}
\end{theoreme}
\selectlanguage{english}
\setcounter{equation}{0}
\section{Introduction}\label{Sec:Intro}
Motivated by the understanding of irreversibility and ``molecular
chaos'' in the Boltzmann equation
\cite{Maxwell1867,Boltzmann1872,Boltzmann1896} Mark Kac proposed in
1956 \cite{kac,kac2} the simpler and seemingly more tractable question
of deriving the \emph{spatially homogeneous} Boltzmann equation from a
\emph{many-particle jump proces}s, and he introduced a rigorous notion of
molecular chaos in this context. He proposed the first proof of the
\emph{propagation of chaos} along time for a simplified collision process for
which series expansions of the solution are available, and he showed
how the many-particle limit rigorously follows from it.
A key motivation for Kac was the microscopic derivation of the
$H$-theorem (monotonicity of the Boltzmann entropy) in this context
which has remained open so far. Kac also raised the natural question
of connecting the asymptotic behavior of the many-particle process and
that of the limit nonlinear equation. In his mind this program was to
be achieved by understanding dissipativity at the level of the linear
many-particle jump process and he insisted on the importance of
estimating its rate of relaxation. This has motivated beautiful works
on this ``Kac's spectral gap problem''
\cite{janvresse,maslen,CarlenCL2003}, but so far this strategy has
proved unsuccessful in obtaining relaxation rates which do not
degenerate in the many-particle limit, see the interesting discussion
in \cite{CCLLV}.
In this Note we present the main results in \cite{mm-article}. In this
paper we develop a quantitative theory of mean-field limit which
\emph{strongly relies on detailed knowledge of the limit nonlinear
equation, rather than on detailed properties of the many-particle
Markov process}. As the main outcome of this theory we prove uniform
in time quantitative propagation of chaos as well as propagation of
entropic chaos, and we prove relaxation rates \emph{independent of the
number of particles} (measured in Wasserstein distance and relative
entropy). All this is done for the two important realistic and
achetypal models of collision, namely hard spheres and true (without
cutoff) Maxwell molecules. This provides a first complete answer to
the questions raised by Kac, however our answer is an ``inverse''
answer in the sense that our methodology is ``top-down'' from the
limit equation to the many-particle system rather than ``bottom-up''
as was proposed by Kac.
\subsection{The Boltzmann equation}
\label{sec:boltzmann-equation}
The {\it spatially homogeneous Boltzmann equation} reads
\begin{equation}\label{el}
\frac{\partial f}{\partial t}(t,v)
= Q(f,f)(t,v), \qquad v \in \mathbb{R}^d, \quad t \geq 0,
\end{equation}
where $d \ge 2$ is the dimension and $Q$ is defined by
\begin{equation*
Q(g,f)(v) = \frac12\,\int _{\mathbb{R}^d \times \mathbb{S}^{d-1}} B(|v-v_*|, \cos \theta)
\left(g'_* f' + g' f_* '- g_* f - g f_* \right) \, dv_* \, d\sigma,
\end{equation*}
where we have used the shorthands $f=f(v)$, $f'=f(v')$, $g_*=g(v_*)$ and
$g'_*=g(v'_*)$. Moreover, $v'$ and $v'_*$ are parametrized by
\begin{equation*
v' = \frac{v+v_*}2 + \frac{|v-v_*|}2 \, \sigma, \qquad
v'_* = \frac{v+v_*}2 - \frac{|v-v_*|}2 \, \sigma, \qquad
\sigma \in \mathbb{S}^{d-1}.
\end{equation*}
Finally, $\theta\in [0,\pi]$ is the deviation angle between $v'-v'_*$
and $v-v_*$ defined by $\cos \theta = \sigma \cdot \hat u$, $u = v-v_*$, $\hat u = u/|u|$,
and $B$ is the Boltzmann collision
kernel determined by physics (related to the cross-section
$\Sigma(v-v_*,\sigma)$ by the formula $B=|v-v_*| \, \Sigma$).
Boltzmann's collision operator has the fundamental properties of
conserving mass, momentum and energy
\begin{equation*}
\frac{d}{dt} \int_{\mathbb{R}^d} f \, \phi(v) \, dv = \int_{\mathbb{R}^d}Q(f,f) \, \phi(v)\,dv = 0, \quad
\phi(v)=1,v,|v|^2,
\end{equation*}
and satisfying the so-called Boltzmann's $H$ theorem which writes (at
the formal level)
\begin{equation*}
- \frac{d}{dt} H(f) := - \frac{d}{dt} \int_{\mathbb{R}^d} f \log f \, dv = -
\frac{d}{dt} H(f |\gamma) := - \frac{d}{dt} \int_{\mathbb{R}^d} f \log
\frac{f}{\gamma} \, dv =
- \int_{\mathbb{R}^d} Q(f,f)\log(f) \, dv \geq 0
\end{equation*}
where $\gamma$ is the gaussian with same mass, momentum and energy as
$f$. Note that the $H$ functional is the \emph{opposite} of the physical
entropy.
We shall consider the folllowing important physical cases for $B$ (see~\cite{mm-article} for more details)
\[
B=\Gamma(|v-v_{*}|) \, b(\cos \theta) \qquad \mbox{ with } \ \Gamma,b \ge
0 \qquad \mbox{ given by one of the following formulas:}
\]
\begin{itemize}
\item[(1)] {\bf (HS)} \textbf{Hard Spheres collision kernel}:
$B(|v-v_*|, \cos \theta)= \Gamma(|v-v_*|) =
C \,
|v-v_*|$ for some $C>0$.
\smallskip
\item[(2)]
{\bf (tMM)} \textbf{True Maxwell Molecules collision kernel}: \\
$B(|v-v_*|, \cos \theta)= b(\cos \theta)
\sim_{\theta \sim 0} C \, \theta^{-5/2}$ for some $C>0$.
\smallskip
\item[(3)] {\bf (GMM)} \textbf{Grad's cutoff Maxwell
Molecules kernel}:
$B(|v-v_*|, \cos \theta)=1.$
\end{itemize}
\subsection{Kac's program}
\label{sec:kacs-program}
{\em Kac's jump process} runs as follows: consider $N$ particles with
velocities $v_1$, \dots, $v_N \in \mathbb{R}^d$. Compute random times for each
pair of particles $(v_i,v_j)$ following an exponential law with
parameter $\Gamma(|v_i-v_j|)$, take the smallest, and perform a
collision $(v_i,v_j) \to (v_i ^*, v_j ^*)$ given by a random choice of
a direction parameter whose law is related to $b(\cos \theta)$, then
recommence. This process can be considered on $\mathbb{R}^{dN}$, however it
leaves invariant some submanifolds of $\mathbb{R}^{dN}$ (depending on the
number of conserved quantities during collision) and can be restricted
to them. In the original simplified model of Kac $d=1$ (scalar
velocities), the direction parameter is $\theta$ with collision rule
$$
v_i ^* = v_i \, \cos \theta + v_j \, \sin \theta, \qquad
v_j ^* = - v_i \, \sin \theta + v_j \, \sin \theta
$$
and the collision process can be restricted to $\mathbb S^{N-1}(\sqrt{
\mathcal E N})$ the sphere with radius $\sqrt{\mathcal E N}$, for
any given value of the \emph{energy} $\mathcal E$. For the more
realistic hard spheres of Maxwell molecules models, $d=3$, the
direction parameter is $\sigma \in \mathbb S^2$ with collision rule
$$
v_i ^* = \frac{v_i + v_j}2 + \frac{|v_i - v_j|}2 \, \sigma, \qquad
v_j ^* = \frac{v_i + v_j}2 - \frac{|v_i - v_j|}2 \, \sigma \qquad
\mbox{ with } \ \sigma \cdot \frac{(v_i-v_j)}{|v_i-v_j|} = \cos \theta
$$
and the collision process can be restricted to the sphere
\begin{equation}\label{sphere-bol}
\mathcal S^N:= \mathbb S^{dN-1}\left(
\sqrt{N \mathcal E}\right) \cap \left\{ v_1 + \dots + v_N =0 \right\}.
\end{equation}
Kac formulated the notion of {\em propagation of chaos} that we shall
now explain. Consider a sequence $(f^N)_{N \ge 1}$ of probabilities on
$\mathbb R^{dN}$: the sequence is said $f$-\textit{chaotic} if $f^N
\sim f^{\otimes N}$ when $N \to \infty$ for some given one-particle
probability $f$ on $\mathbb R^d$. The meaning of this convergence is the
following: convergence in the weak measure topology for any marginal
depending on a finite number of variables. This is a \emph{low
correlation} assumption. It was clear since Boltzmann that in the
case when the joint probability density $f^N$ of the $N$-particle
system is tensorized \emph{during some time interval} into $N$ copies
$f^{\otimes N}$ of a $1$-particle probability density, then the latter
would satisfy the limit nonlinear Boltzmann equation during this time
interval. In general interactions between particles prevent any
possibility of propagation of the ``tensorization'' property, however
if the weaker property of chaoticity can be propagated along time in
the correct scaling limit it is sufficient for deriving the limit
equation.
Kac hence proved the propagation of chaos (with no rate) on the
simplified collision rule above (with $B=1$). His beautiful
combinatorial argument is based on an infinite series ``tree''
representation of the solution according to the collision history of
particles, and a Leibniz derivation-like formula for the iterated
$N$-particle operator acting on tensor products.
He then raises several questions that we schematize as follows:
\begin{enumerate}
\item The first one is concerned with the restriction of the models as
compared to realistic collision processes: {\bf can one prove
propagation of chaos for the hard spheres collision process?}
\item Following closely the spirit of the previous question
it seems to us
very natural to ask whether {\bf one can prove propagation of chaos
for the true Maxwell molecules collision process}?
This is related with \emph{long-range interactions} and
\emph{fractional derivative operators}.
\item
Kac {\bf conjectures the propagation of the convergence of the
$N$-particle $H$-functional towards the limit $H$-functional of
the solution to the limit equation along time in the mean-field
limit}. Since the latter always decays for a many-particle jump
process, in his words ``{\it If the above steps could be made
rigorous we would have a thoroughly satisfactory justification of
Boltzmann's $H$-theorem.}''
\item He finally discusses the relaxation times, with the goal of
deriving relaxation times of the limit equation from the
many-particle system. This imposes to have estimates
\emph{independent of the number of particles} on this relaxation
times:
{\bf can one
prove relaxation times {\it independent of the number of
particles} in Wasserstein distance and/or
relative entropy?}
\end{enumerate}
This paper is concerned with solving the four questions
outlined above.
\section{Main results}\label{Sec:Main}
\subsection{A few words on previous results}
\label{sec:review}
For Boltzmann collision processes, Kac \cite{kac} has proved the
propagation of chaos in the case of his baby one-dimensional model. It
was generalized by McKean \cite{McKean1967} to the Boltzmann collision
operator for ``Maxwell molecules with cutoff'', i.e. the case {\bf
(GMM)} above (see also \cite{T2} for a partial result for non-cutoff
Maxwell molecules). Gr\"unbaum~\cite{Grunbaum} then proposed in a very
compact and abstract paper another method for dealing with hard
spheres, based on the Trotter-Kato formula for semigroups and a clever
functional framework. Unfortunately this paper was incomplete for
several reasons (see the discussion in \cite{mm-article}). A
completely different approach was undertaken by Sznitman in the
eighties \cite{S1,S6} and he gave a full proof of propagation of chaos
for hard spheres by a probabilistic (non-constructive) approach. Let
us also emphasize several quantitative results on a finite time
interval by Graham, M\'el\'eard and Fournier for Maxwell molecules
models \cite{GM,FM7,FM10}, and the works on the so-called \leavevmode\raise.3ex\hbox{$\scriptscriptstyle\langle\!\langle$~} Kac's
spectral gap problem\leavevmode\raise.3ex\hbox{~$\!\scriptscriptstyle\,\rangle\!\rangle$}{} \cite{janvresse,maslen,CarlenCL2003,CCLLV}.
\subsection{Main results}
\label{sec:intromainresults}
\begin{theorem}[Summary of the main results]\label{theo:main}
Consider some initial distribution $f_0 \in P(\mathbb{R}^d) \cap L^\infty$
with compact support or polynomial moment bounds, taken to be
centered without loss of generality. Consider the corresponding
solution $f_t$ to the spatially homogeneous Boltzmann equation for
hard spheres of Maxwell molecules, and the solution $f^N_t$ of the
corresponding Kac's jump process starting either (a) from the
tensorization $f_0 ^{\otimes N}$ of $f_0$ or (b) the latter
conditionned to $\mathcal S^N$ (defined in (\ref{sphere-bol})).
The results in \cite{mm-article} can be classified into three main statements:
\begin{enumerate}
\item {\bf Quantitative uniform in time propagation of chaos (with
any number of marginals)}:
\begin{equation*
\forall \, N \ge 1, \ \forall \, 1 \le \ell \le N, \quad
\sup_{t \ge 0} {W_1 \left( \Pi_{\ell} f^N
_t, \left( f_t ^{\otimes \ell} \right) \right) \over \ell} \le \alpha(N)
\end{equation*}
for some $\alpha(N) \to 0$ as $N \to \infty$, where $\Pi_\ell g^N$
stands for the $\ell$-marginal of an $N$-particle distribution $g^N$,
and where $W_1$ is the Wasserstein distance between probabilities on
$\mathbb R^{d\ell}$:
$$
W_1(p_1,p_2) := \sup_{[ \varphi ]_{\mbox{{\tiny {\em Lip}}}(\mathbb
R^{d\ell})} \le 1} \int_{\mathbb R^{d\ell}} \varphi \, (dp_1 - dp_2)
\qquad \mbox{where } \ [\cdot]_{\mbox{{\tiny {\em Lip}}}} \ \mbox{
denotes the Lipschitz semi-norm.}
$$
In the case (a) $f^N _0 = f^{\otimes N} _0$ one has moreover explicit
power law rate (for Maxwell molecules) or logarithmic rate (for hard
spheres) estimates on $\alpha$.
\item {\bf Propagation of entropic chaos}: Consider the case (b) where the
initial datum of the many-particle system is restricted to $\mathcal S^N$.
Then if the initial datum is entropy-chaotic in the sense
$$
\frac1N \, H\left( f^N _0 | \gamma^N \right) \to H\left(f_0 | \gamma
\right), \quad N \to +\infty
$$
\begin{equation}\label{rel-entropie}
\mbox{with} \quad H\left( f^N _0 | \gamma^N \right) := \int_{\mathcal
S^N} \frac{df^N _0}{d \gamma^N} \, \log \frac{df^N
_0}{d\gamma^N} \, \gamma^N(dV) \ \mbox{ and } \ H\left(f_0 | \gamma \right) :=
\int_{\mathbb{R}^d} f_0 \, \log \frac{f_0}{\gamma} \, dv
\end{equation}
and where $\gamma$ is the gaussian equilibrium with energy $\mathcal
E$ and $\gamma^N$ is the uniform probability measure on $\mathcal
S^N$, then the solution is also entropy-chaotic for any later time:
$$
\forall \, t \ge 0, \quad \frac1N \, H\left( f^N _t | \gamma^N \right)
\to H\left(f_t | \gamma \right), \quad N \to +\infty.
$$
Since our $f_0 ^N$ is entropy-chaotic, this proves the
derivation of the $H$-theorem in this context.
\item {\bf Quantitative estimates on relaxation times, independent of
the number of particles}: Consider the case (b) where the
initial datum of the many-particle system is restricted to $\mathcal
S^N$. Then
\begin{equation*
\forall \, N \ge 1, \ \forall \, 1 \le \ell \le N, \ \forall \, t
\ge 0, \quad
{W_1 \left( \Pi_{\ell} f^N
_t, \Pi_\ell \left( \gamma^N \right) \right) \over \ell} \le
\beta(t) \qquad \mbox{with} \ \beta(t) \to 0, \ t \to 0.
\end{equation*}
Moreover in the case of Maxwell molecules, and assuming moreover that
the Fisher information of the initial datum $f_0$ is finite:
\begin{equation}\label{fisher}
\int_{\mathbb{R}^d} \frac{\left| \nabla_v f_0 \right|^2}{f_0} \, dv <
+\infty,
\end{equation}
the following estimate also holds:
$
\forall \, N \ge 1, \quad 0 \le \frac1N \, H\left( f^N _t | \gamma^N \right)
\le \beta(t) \quad \mbox{with} \ \beta(t) \to 0, \ t \to 0.
$
\end{enumerate}
\end{theorem}
\section{A few words on the methods and proofs}\label{Sec:Proofs}
Let us briefly explain some ideas underlying the result
Theorem~\ref{theo:main}-(i). The other results are then obtained on
the basis of this key estimate, combined with the other latest
results obtained in this field.
\begin{itemize}
\item We aim at reducing the problem to a stability analysis when
approximating a linear semigroup. To this purpose a key idea is to
compare the linear $N$-particle dual evolution in $C_b(\mathbb
R^{dN})$ with the (linear!) \emph{push-forward} evolution associated
with the limit equation.
\item This push-forward semigroup is defined as follows: if $S^{N
\!L}_t$ denotes the nonlinear semigroup of \eqref{el}, this
push-forward semigroup is defined on $P(P(\mathbb R^d))$ by
$T^\infty _t[\Phi](f) = \Phi(S^{N \!L}_t(f))$.
\item In order to make this comparison between semigroups, we use the
empirical measure $\mu^N _V = (\sum_{i=1} ^N \delta_{v_i})/N$ in
order to embed the dynamics in $P(\mathbb R^{dN})$ into a dynamics
in $P(P(\mathbb R^d))$.
\item One then considers the following term to be estimated
$$
\left| \left \langle \left( S^N_t(f_0^{N}) - \left( S _t ^{NL}
(f_0)
\right)^{\otimes N} \right), \varphi
\otimes 1^{\otimes N-\ell} \right\rangle \right|
$$
for some test function $\varphi$ only depending on $\ell$ variables.
\item The approximation of these marginals by empirical
measure estimates yields a first error term on the $N$-particle
semigroup
$$
\left| \left\langle S^N_t(f_0^N),
\varphi \otimes
1^{\otimes N-\ell} \right\rangle -
\left \langle S^N_t(f_0^N),
R^\ell_\varphi \circ \mu^N_V \right\rangle \right| \qquad
\mbox{ with } \ \ R^\ell _\varphi (f) := \int_{\mathcal R^{d\ell}}
\varphi \, f^{\otimes \ell}(dv_1 \dots dv_\ell)
$$
which is controlled by combinatorial arguments, and then a second
error term
$$
\left| \left\langle f_0^N,
(T_t ^\infty R^\ell_\varphi ) \circ \mu^N_V) \right\rangle
- \left\langle (S _t ^{NL} (f_0))^{\otimes \ell} ,
\varphi \right\rangle \right|
$$
which is controlled thanks a stability for measure solutions of the
limit equation.
\item Finally there remains the most important term where the two
dynamics are effectively compared
$$
\left| \left\langle f_0^N, T^N_t ( R^\ell_\varphi \circ
\mu^N_V) \right\rangle
- \left\langle f_0^N,
(T_t ^\infty R^\ell_\varphi ) \circ \mu^N_V) \right\rangle
\right|.
$$
This term is controlled by using (1) a quantitative argument \`a la
Trotter-Kato in order to express the difference of semigroups in terms
of the difference of their generators $G^N$ and $G^\infty$, (2) a
\emph{consistency estimate} between those generators, (3) a \emph{stability
estimate} on the limit equation.
\item There is a \emph{loss of derivative} in the consistency estimate
in the sense of differentiable functions acting on $P(\mathbb R^d)$,
which lead us to develop a differential calculus on this space
adapted to our purpose.
\item The stability estimate means, once translated on the original
nonlinear semigroup $S^{N \!L}_t$ of \eqref{el}, a propagation of a
bound $C^{1+\theta}(P(\mathbb R^d))$ on $S^{N \!L}_t$. The role
played by such stability estimates is a key novelty of our
study. Proving them for Boltzmann is also one of the most technical
aspects of \cite{mm-article}.
\item There are many possible choices of distances on the space of
probabilities (total variation but also many non-equivalent weak
measure distances), and it is a crucial point that our method is
flexible enough to allow for many such different choices adapted to
the equations it is applied to.
\end{itemize}
|
\section{#1}\setcounter{equation}{0}}
\numberwithin{equation}{section}
\newtheorem{theorem}{Theorem}[section]
\mynewtheorem[theorem]{lemma}{Lemma}{Lemmas}
\mynewtheorem[theorem]{proposition}{Proposition}{Propositions}
\mynewtheorem[theorem]{corollary}{Corollary}{Corollaries}
\theorembodyfont{\normalfont}
\mynewtheorem[theorem]{defi}{Definition}{Definitions}
\mynewtheorem[theorem]{conjecture}{Conjecture}{Conjectures}
\mynewtheorem[theorem]{example}{Example}{Examples}
\mynewtheorem[theorem]{notation}{Notation}{Notations}
\mynewtheorem[theorem]{remark}{Remark}{Remarks}
\mynewtheorem[theorem]{remarks}{Remarks}{Remarks}
\theoremsymbol{\ensuremath{\Box}}
\theoremseparator{:}
\newtheorem*{proof}{Proof}
\newcommand{\begin{equation}}{\begin{equation}}
\newcommand{\end{equation}}{\end{equation}}
\newcommand{\Leq}[1]{\label{#1}\end{equation}}
\newcommand{\emptyset}{\emptyset}
\renewcommand {\l}{\left}
\newcommand {\ri}{\right}
\newcommand {\vep}{\varepsilon
\newcommand {\eps}{\epsilon}
\newcommand {\vv}{\varphi}
\newcommand {\Ll}{\Lambda_{l}}
\newcommand {\LA}{\left\langle}
\newcommand {\RA}{\right\rangle}
\newcommand {\pa}{\partial}
\newcommand {\pih}{\hat{\pi}}
\newcommand {\eh}{{\textstyle \frac{1}{2}}}
\newcommand {\ed}{{\textstyle \frac{1}{3}}}
\newcommand {\ev}{{\textstyle \frac{1}{4}}}
\newcommand {\dv}{{\textstyle \frac{3}{4}}}
\newcommand {\ea}{{\textstyle \frac{1}{8}}}
\renewcommand {\dh}{{\textstyle \frac{3}{2}}}
\newcommand {\inInd}{ \in \{1,\ldots,n\} }
\newcommand {\ar}{\rightarrow}
\newcommand {\sign}{\mathrm{sign}}
\newcommand {\const}{\mathrm{const}}
\newcommand {\ess}{\operatorname{ess}}
\newcommand {\Euclid}{\mathrm{Euclid}}
\newcommand {\id}{\mathrm{id}}
\newcommand {\Id}{\mathrm{Id}}
\newcommand {\Span}{\mathrm{span}}
\renewcommand {\Im}{\mathrm{Im}}
\newcommand {\qm}{\mathrm{qm}}
\newcommand {\op}{\mathrm{op}}
\renewcommand {\Re}{\mathrm{Re}}
\newcommand {\rot}{\mathrm{rot}}
\newcommand {{\hat{\rho}\,}}{\mathrm{tr}}
\newcommand {\bi}{\mathbf{i}_X}
\newcommand {\bfi}{\mathbf{i}}
\newcommand {\bC}{{\mathbb C}}
\newcommand {\bD}{{\mathbb D}}
\newcommand {\bE}{{\mathbb E}}
\newcommand {\bN}{{\mathbb N}}
\newcommand {\bR}{{\mathbb R}}
\newcommand {\bZ}{{\mathbb Z}}
\newcommand {\bT}{{\mathbb T}}
\newcommand {\bP}{{\mathbb P}}
\newcommand {\bQ}{{\mathbb Q}}
\newcommand {\bV}{{\mathbb V}}
\newcommand {\ImH}{{\mathrm{Im}\bH}}
\newcommand{{\upharpoonright}}{{\upharpoonright}}
\newcommand{\idty}{{\rm 1\mskip-4mu l}}
\newcommand{\vidty}{{\vec{\idty}}}
\newcommand{\cA}{{\cal A}}
\newcommand{\cB}{{\cal B}}
\newcommand{\cC}{{\cal C}}
\newcommand{\cD}{{\cal D}} %
\newcommand{\cE}{{\cal E}} %
\newcommand{\cF}{{\cal F}}
\newcommand{\cG}{{\cal G}}
\newcommand{{\cal H}}{{\cal H}}
\newcommand{\cI}{{\cal I}}
\newcommand{{\cal J}}{{\cal J}}
\newcommand{{\cal K}}{{\cal K}}
\newcommand{{\mathcal L}}{{\mathcal L}}
\newcommand{{\cal M}}{{\cal M}}
\newcommand{{\cal N}}{{\cal N}}
\newcommand{\cO}{{\cal O}}
\newcommand{{\cal P}}{{\cal P}}
\newcommand{\cS}{{\cal S}}
\newcommand{\cV}{{\cal V}} %
\newcommand{\cW}{{\cal W}}
\newcommand{{\cal N\!P}}{{\cal N\!P}}
\newcommand{{\cal T}}{{\cal T}}
\newcommand{{\cal X}}{{\cal X}}
\newcommand{\overline}{\overline}
\newcommand{\l(\! \begin{array}}{\l(\! \begin{array}}
\newcommand{\end{array}\!\ri)}{\end{array}\!\ri)}
\newcommand{\bsm}{\left(\begin{smallmatrix}}
\newcommand{\esm}{\end{smallmatrix}\right)}
\newcommand{\nonumber}{\nonumber}
\newcommand{T\!M}{T\!M}
\newcommand{\hat{A}}{\hat{A}}
\newcommand{\hat{\Sigma}}{\hat{\Sigma}}
\newcommand{{\hat{\Sigma}^*}}{{\hat{\Sigma}^*}}
\newcommand{{\vec{p}\,}}{{\vec{p}\,}}
\newcommand{{\vec{q}\,}}{{\vec{q}\,}}
\renewcommand{\v}{{\vec{v}\,}}
\newcommand{{\hat{s}\,}}{{\hat{s}\,}}
\newcommand{{\hat{q}\,}}{{\hat{q}\,}}
\newcommand{{\hat{L}\,}}{{\hat{L}\,}}
\newcommand{{\hat{O}\,}}{{\hat{O}\,}}
\newcommand{{\hat{U}}}{{\hat{U}}}
\newcommand{{\hat{\mu}}}{{\hat{\mu}}}
\newcommand{\tlambda}{{\hat{\lambda}}}
\newcommand{{\hat{P}}}{{\hat{P}}}
\newcommand{{\hat{p}}}{{\hat{p}}}
\newcommand{{\hat{\pi}}}{{\hat{\pi}}}
\newcommand{{\hat{B}}}{{\hat{B}}}
\newcommand{{\tilde{B}}}{{\tilde{B}}}
\newcommand{{\hat{H}}}{{\hat{H}}}
\newcommand{\tK}{{\hat{K}}}
\newcommand{\tM}{{\hat{M}}}
\newcommand{\tV}{{\hat{V}}}
\newcommand{\tW}{{\hat{W}}}
\newcommand{{\hat{\Phi}}}{{\hat{\Phi}}}
\newcommand{{\hat{z}}}{{\hat{z}}}
\newcommand{{\overline{v}}}{{\overline{v}}}
\newcommand{\qtext}[1]{\quad\text{#1}\quad}
\newcommand{\qtextq}[1]{\quad\text{#1}\quad}
\RequirePackage{ifthen}
\newcommand{\abs}[2][]{\lvert#2\rvert\ifthenelse{\equal{#1}{}}{}{_{#1}}}
\newcommand{\bigabs}[2][]{\bigl|#2\bigr|\ifthenelse{\equal{#1}{}}{}{_{#1}}}
\newcommand{\Bigabs}[2][]{\Bigl|#2\Bigr|\ifthenelse{\equal{#1}{}}{}{_{#1}}}
\newcommand{\norm}[2][]{\lVert#2\rVert\ifthenelse{\equal{#1}{}}{}{_{#1}}}
\newcommand{\setsize}[2][]{\lvert#2\rvert\ifthenelse{\equal{#1}{}}{}{_{#1}}}
\newcommand{\bra}[1]{\langle#1\rvert}
\newcommand{\ket}[1]{\lvert#1\rangle}
\newcommand{\bk}[2]{\langle#1\,,\,#2\rangle}
\newcommand{\bkop}[3]{\langle#1,#2 #3\rangle}
\newcommand{\kb}[2]{\lvert#1\rangle\langle#2\rvert}
\newcommand{{\dot{x}}}{{\dot{x}}}
\newcommand{\mathrm{dist\,}}{\mathrm{dist\,}}
\newcommand{\mathrm{supp}}{\mathrm{supp}}
\newcommand{\mathrm{diam}}{\mathrm{diam}}
\newcommand{\northeast}
{\raisebox{1.5mm}{$\lrcorner$}\raisebox{-2mm}{$\ulcorner$}}
\newcommand{\northwest}
{\raisebox{-2mm}{$\urcorner$}\raisebox{1.5mm}{$\llcorner$}}
\newcommand{\,\frown\hspace*{-2.7mm}\mid\ }{\,\frown\hspace*{-2.7mm}\mid\ }
\newcommand{\note}[1]{\marginpar{\raggedright\small#1}}
\newcommand{\red}[1]{\textcolor{red}{#1}}
\newcommand{\blue}[1]{\textcolor{blue}{#1}}
\newcommand{\green}[1]{\textcolor{green}{#1}}
\newcommand{\cs}[1]{\textcolor{cyan}{#1}}
\providecommand{\Vintern}[1]{\left(\begin{smallmatrix}#1\end{smallmatrix}\right)}%
\providecommand{\V}[1]{{\mathchoice{\begin{pmatrix}#1\end{pmatrix}}{\Vintern{#1}}{\Vintern{#1}}{\Vintern{#1}}}
\providecommand{\xto}{\xrightarrow}
\providecommand{\Union}{\bigcup}
\providecommand{\union}{\cup}
\providecommand{\Isect}{\bigcap}
\providecommand{\isect}{\cap}
\providecommand{\interior}{\operatorname{int}}
\providecommand{\ve}{\varepsilon}
\providecommand{\Mo}{\mathbf M_\omega}
\providecommand{\textq}[1]{\text{#1}\quad}
\providecommand{\dnto}{\searrow}
\providecommand{\ceil}[1]{\lceil#1\rceil}
\providecommand{\floor}[1]{\lfloor#1\rfloor}
\begin{document}
\title {Classical Motion in Random Potentials}
\author{Andreas Knauf
\thanks{Department Mathematik,
Universit\"at Erlangen-N\"urnberg,
Cauerstr.~11, D--91\,058 Erlangen, Germany.
e-mail: <EMAIL>}
\and Christoph Schumacher
\thanks{Fakult\"at f\"ur Mathematik,
Technische Universit\"at Chemnitz,
Reichenhainerstr.~$41$, D--09\,126 Chemnitz, Germany.
e-mail: <EMAIL>}
}
\date{October 2011}
\maketitle
\begin{abstract}
We consider the motion of a classical particle under the influence
of a random potential on~$\bR^d$, in particular the distribution of
asymptotic velocities and the question of ergodicity of time evolution.
\end{abstract}
\tableofcontents
\section{Introduction}\label{sec1}
Since its introduction to physics by Einstein
and its mathematical foundation by Wiener, Brownian motion is
considered one cornerstone of probability and of thermodynamics.
It thus may seem natural to expect
that classical motion in a spatially
homogeneous force field leads in the large time limit
to deterministic diffusion.
Such results can in fact be proven for a periodic Lorentz gas
with finite horizon (related to the Sinai billiard)
or for coulombic periodic potentials,
see \cite{BS81} respectively \cite{Kna87, DL91}.
However, in both cases the dynamics is non--smooth, of billiard
type for the Lorentz gas and with orbits locally approximating
Keplerian conic sections for the second case.
We show here (\cref{theo}), that the
motion in a bounded smooth potential
is incompatible with uniform hyperbolicity.
If periodic scatterers can lead to diffusive motion,
the more this should be true for random scatterers.
However, this is certainly not the case for 1D and
also wrong for more degrees of freedom and Poisson potentials
(\cref{sec5,sec5a}).
In this and other cases the Hamiltonian flow is not even ergodic
(\cref{thmPoisson}).
For the random Coulomb case, however, we show in
\cref{thm:toptrans} that the flow is
typically topologically transitive for large energies.
Related results on the motion in random
configurations of convex scatterers have been derived by
Marco Lenci and collaborators.
In \cite{Len03}, the planar situation with a finite modification
of periodic scatterers was studied.
In \cite{CLS10} (see also the references of that article)
recurrence for particles in quenched tubes
with random scatterers has been proven.
The literature on the corresponding quantum problem of
Schr\"odinger operators with random potentials is much broader.
See \emph{e.g.}, \cite{LMW03}, and \cite{Ve08} with its extensive
references.
In the article \cite{DRF11} quantum diffusion in a thermal
medium has been proven.
We describe the structure of the paper.
Random potentials arise in different guises.
The one studied most extensively is based on a regular lattice~${\mathcal L}$
in configuration space. If $J$ indexes the single site potentials,
on assumes a measure on the space $J^{\mathcal L}$,
for which the ${\mathcal L}$--action is ergodic.
This is studied in \cref{sec2}.
One result is an almost deterministic distribution of
asymptotic velocities (\cref{sec3}).
One problem concerning the Liouville measure
on the energy surface is discussed in \cref{sec4a}.
Depending on the exact exponential decay rate,
the set of singular energy values may or may not be typically dense.
Poissonian random potentials are studied in \cref{sec4,sec5a}.
Relationships between different notions of ergodicity
and some of their dynamical implications are discussed in \cref{sec5}.
\cref{sec6} concerns deterministic potentials, and the
geometry compatible with a uniformly hyperbolic structure.
Finally, we treat random coulombic potentials in \cref{sec9,sect10},
showing topological transitivity and ergodicity of the compactified
flow for energy surfaces of large energy.
\bigskip
\noindent
{\bf Acknowledgement:} We thank Boris Gutkin for the useful conversations.
\section{The Lattice Case}\label{sec2}
We assume the random potential to be based on \emph{short range} single site
potentials $W_j\in C^\eta(\bR^d,\bR)$,
\nomenclature[AW]{$W_j$}{single site potential}{}{}%
\nomenclature[Ad]{$d$}{dimension}{}{}%
indexed by $j\in J$, $\setsize J\in\bN$, $\setsize J\ge2$%
\nomenclature[AJ]{$J$}{index set for single site potentials}{}{}.
Namely $\eta\in\bN\cup\{\infty\}$, $\eta\ge2$%
\nomenclature[Geta]{$\eta$}{order of differentiability}{}{}
and
\begin{equation}
\abs{\pa^\alpha W_j(q)}\leq\frac{C_\alpha}{\LA q\RA^{d+\vep}} \qquad
\bigl(q\in\bR^d,\ \alpha\in\bN_0^d,\ \abs\alpha\leq\eta\bigr)
\Leq{as:sisi}
\nomenclature[ACalpha]{$C_\alpha$}{constant}{}{}%
\nomenclature[Galpha]{$\alpha$}{multi index}{}{}%
with $\LA q\RA:=\sqrt{1+\norm{q}^2}$, for constants $C_\alpha>0$.
The single site potentials are placed on a regular lattice ${\mathcal L}\subseteq\bR^d$%
\nomenclature[AL]{${\mathcal L}$}{regular lattice in $\bR^d$}{}{}
with basis $\ell_1,\dotsc,\ell_d\in\bR^d$%
\nomenclature[Aell1]{$\ell_1,\dotsc,\ell_d$}{basis of ${\mathcal L}$}{}{}
according to $\omega\in\Omega:=J^{\mathcal L}$%
\nomenclature[GOmega]{$\Omega$}{configurations of single site potentials}{}{}%
\nomenclature[Gomega]{$\omega$}{configuration of single site potentials}{}{}
to give the random potential
\[V\colon M:=\Omega\times\bR^d\to\bR \qtextq{,} V(\omega,q):=\sum_{\ell\in{\mathcal L}}
W_{\omega(\ell)}(q-\ell)\]
\nomenclature[AV]{$V$}{(random) potential}{}{}%
\nomenclature[Al]{$\ell$}{lattice vector}{}{}%
on \emph{extended configuration space}~$M$%
\nomenclature[AM]{$M$}{extended configuration space}{}{}.
We use the discrete topology on~$J$ and product topology on~$\Omega$.
The probability measure~$\beta$%
\nomenclature[Gbeta]{$\beta$}{probability measure on $\Omega$}{}{}
on $\bigl(\Omega,\cB(\Omega)\bigr)$
is assumed to be invariant w.r.t.\ the action
\begin{equation}\label{eq:vartheta}
\vartheta\colon{\mathcal L}\times\Omega\to\Omega\qtextq{,}
(\ell,\omega)\mapsto\vartheta_\ell(\omega)\qtextq{with}
\vartheta_\ell(\omega)(\ell'):=\omega(\ell'+\ell)\text.
\end{equation}
\nomenclature[Gtheta]{$\vartheta$}{${\mathcal L}$--action on $\Omega$}{}{}%
Unless we explicitly say the contrary, $\beta$ is assumed to be
$\vartheta$--ergodic (a simple example being a product measure
$\beta=\bigotimes_{\ell\in{\mathcal L}}\hat{\beta}$
with a probability measure~$\hat{\beta}$%
\nomenclature[Gbetahat]{$\hat\beta$}{probability measure on~$J$}{}{}
on~$J$).
Ergodicity and $\setsize J\ge2$ imply that~$\beta$ is non--atomic.
The short range conditions above imply that for all $\omega\in\Omega$
the potentials
\[V_\omega\colon\bR^d\to\bR\qtextq{,} V_\omega(q):=V(\omega,q)\]
are as smooth as the single site potentials $W_j$.
Moreover~$V$ itself is continuous and bounded,
together with its partial derivatives $\pa^\alpha V,\ \abs\alpha\leq\eta$.
\begin{remark}\label{rem:A}
More relevant than the \emph{range}
\begin{equation*}
[V_{\min},V_{\max}]:=\overline{V(M)}
\end{equation*}
\nomenclature[AVmin]{$V_{\min}$}{infimum of~$V$}{}{}%
\nomenclature[AVmax]{$V_{\max}$}{supremum of~$V$}{}{}%
of the potential is its \emph{essential range}
($\lambda^d$%
\nomenclature[Glambdad]{$\lambda^d$}{Lebesgue measure on $\bR^d$}{}{}
denoting Lebesgue measure on $\bR^d$):
\begin{equation*}
\mathrm{supp}\bigl(V(\beta\otimes\lambda^d)\bigr)
=[V_{\ess\min},V_{\ess\max}]\text.
\end{equation*}
\nomenclature[AVessmin]{$V_{\ess\min}$}{essential infimum of~$V$}{}{}%
\nomenclature[AVessmax]{$V_{\ess\max}$}{essential supremum of~$V$}{}{}%
By $\vartheta$--ergodicity of $\beta$, this is a deterministic set:
\[\mathrm{supp} \bigl(V_\omega(\lambda^d)\bigr)
= \mathrm{supp}\bigl(V(\beta\otimes\lambda^d)\bigr) \qquad
(\beta\text{--a.s.})\text.\]
As~$V$ is uniformly bounded together with its first derivatives, this is in fact
a bounded interval:
$\mathrm{supp}\bigl(V_\omega(\lambda^d)\bigr)=\overline{V_\omega(\bR^d)}$.
In general the essential range of~$V$ is a proper subinterval of its range.
\end{remark}
Thus the flow $\Phi\colon\bR\times P\to P$%
\nomenclature[GPhi]{$\Phi$}{hamiltonian flow}{}{}
generated by the Hamiltonian function
\[H\colon P\to\bR \qtextq{,} H(\omega,p,q):=\eh\norm p^2+V(\omega,q)\]
\nomenclature[AH]{$H$}{hamiltonian function}{}{}%
on \emph{extended phase space}
\[P:=\Omega\times\bR^d\times\bR^d\]
\nomenclature[AP]{$P$}{extended phase space}{}{}%
uniquely exists for all times.
We write
\[\Phi^t\colon P\to P\text{ , } \Phi^t(\omega,p_0,q_0):=\Phi(t,\omega,p_0,q_0)=
\l(\omega\, ,\, p^t(\omega,p_0,q_0)\, ,\, q^t(\omega,p_0,q_0)\ri)\]
\nomenclature[Apt]{$p^t$}{momentum at time~$t$}{}{}%
\nomenclature[Aqt]{$q^t$}{location at time~$t$}{}{}%
for the solution of the initial value problem at time $t\in\bR$%
\nomenclature[At]{$t$}{time}{}{}.
Whenever a fixed $\omega\in\Omega$ is considered, we write it as a subscript
(\emph{e.g.}\ $\Phi_\omega^t\equiv(p_\omega^t,q_\omega^t)\colon\bR^{2d}\to\bR^{2d})$.
See
\cref{abb:realisation}
for a realisation of $t\mapsto q_\omega(t,x_0)$, with lattice ${\mathcal L}=\bZ^2$.
\begin{figure}[h
\begin{center}
\ifpdf
\includegraphics[viewport=90 375 440 715,width=9cm,clip]{zufall4.pdf}
\else
\includegraphics[viewport=90 375 440 715,width=9cm,clip]{zufall4.ps}
\fi
\end{center}
\caption{Path in configuration space}
\label{abb:realisation}
\end{figure}
The space~$P$ is equipped with the locally finite Borel measure
$\mu:=\beta\otimes\lambda^{2d}$%
\nomenclature[Gmu]{$\mu$}{invariant measure on~$P$}{}{}.
The Hamiltonian flow~$\Phi$ leaves~$\mu$ invariant.
The lattice acts on extended phase space via the continuous group action
\[\Theta\colon{\mathcal L}\times P\to P \qtextq{,}(\ell,\omega,p,q)\,\mapsto\,
\Theta_\ell(\omega,p,q) := \bigl(\vartheta_\ell(\omega),p,q-\ell\bigr)\]
\nomenclature[GTheta]{$\Theta$}{${\mathcal L}$--action on~$P$}{}{}%
and leaves $\mu$ invariant:
\begin{equation}
\mu\circ\Theta_\ell=\mu \qquad (\ell\in{\mathcal L})\text.
\Leq{A}
Similarly for all lattice vectors $\ell\in{\mathcal L}$
\begin{equation}
H\circ\Theta_\ell=H \qtextq{and thus} \Theta_\ell\circ\Phi^t=\Phi^t\circ
\Theta_\ell \qquad (t\in\bR)\text.
\Leq{B}
By the Hamiltonian character of the motion the energy is invariant under
the time evolution:
\begin{equation}
H\circ\Phi^t=H \qquad (t\in\bR)\text.
\Leq{C}
Thus we obtain a one--parameter family of Borel measures $\mu_E$%
\nomenclature[GmuE]{$\mu_E$}{invariant measure on $H^{-1}(-\infty,E]$}{}{}
on~$P$, given by
\begin{equation}
\mu_E(B):=\mu \l(B\cap H^{-1}(-\infty,E])\ri)
\qquad \bigl(B\in\cB(P)\bigr)\text,
\Leq{mu:E}
parametrised by the energy $E\in\bR$%
\nomenclature[AE]{$E$}{energy}{}{}.
By \eqref{A}, \eqref{B} and \eqref{C}
these measures~$\mu_E$ are $\Theta$--~and $\Phi$--invariant, too.
A relevant quantity is \emph{asymptotic} velocity
\begin{equation}
{\overline{v}}^\pm\colon P\to\bR^d \qtextq{,} {\overline{v}}^\pm(\omega,x_0):=\lim_{T\to\pm\infty}
\frac{q_\omega(T,x_0)}{T}\text.
\Leq{D}
\nomenclature[Avbar]{${\overline{v}}$}{asymptotic velocity}{}{}%
\begin{remark}[Non--Existence of Asymptotic Velocity] \label{rem:never}
For Hamiltonian motion in bounded potentials the asymptotic velocity may not exist for \emph{any} initial condition
$x_0\in H^{-1}(E)$ for \emph{all} energies $E$ above some threshold.
This is the case for centrally symmetric potentials
($V(q)=\tilde V(\norm q)$), with $\tilde V$%
\nomenclature[AVtilde]{$\tilde V$}{slowly varying function}{}{}
being periodic in a slowly varying function like, for example,
$\tilde V(r)=\cos(\log(r+1))$.
It is also possible to construct examples of truly random potentials
where for some $\omega\in\Omega$ these limits do not exist on
$ H_\omega^{-1}(E)$ for any $E>E_0$:
The lattice ${\mathcal L}:=\bZ^d$ with fundamental domain $\cD=[0,1]^d$%
\nomenclature[AD]{$\cD$}{fundamental domain of~${\mathcal L}$}{}{}
admits an adapted partition of unity $F_\ell$%
\nomenclature[AFell]{$F_\ell$}{partition of unity}{}{},
$(\ell\in{\mathcal L})$
\[{\textstyle \sum_{\ell\in{\mathcal L}}} F_\ell(q) = 1 \qquad (q\in\bR^d)\]
with $F_\ell(q) := F(q-\ell)$
for
$F:=\idty_\cD\,*\,f$ with some
$f\in C_c^\infty\bigl(\bR^d,[0,\infty)\bigr)$ and $\int_{\bR^d}f\,dx=1$.
\nomenclature[Aidentity]{$\idty_\cD$}{indicator function of~$\cD$}{}{}%
\nomenclature[Af]{$f$}{smooth nonnegative function}{}{}%
We now take $J:=\{0,1,\ldots,d\}$ as index set for the single site potentials
$W_j:=j\,F$.
We choose $\omega\in\Omega=J^{(\bZ^d)}$ of the form
$\omega(\ell)=\sum_{k=1}^{d} \widetilde\omega_k(\ell_k)$%
\nomenclature[Gomegatildek]{$\widetilde\omega_k$}{slowly varying configuration}{}{},
with $\widetilde\omega_k:\bZ\to \{0,1\}$ slowly varying like above.
Then the motion in the realization $V_\omega$
is separable in cartesian coordinates,
and thus for no $E>E_0:=d$%
\nomenclature[AEzero]{$E_0$}{energy threshold}{}{}
and initial condition $x_0\in H_\omega^{-1}(E)$
the asymptotic velocities ${\overline{v}}^\pm(x_0)$ exist.
By allowing for arbitrary modifications of $\omega$ over finite subsets
of~$\bZ^d$ the set of these $\widehat{\omega}$%
\nomenclature[Gomegahat]{$\widehat\omega$}{finite modification of $\omega$}{}{}
is even dense in $\Omega$,
and asymptotic velocity does not exist for any initial condition
$x_0\in H_{\widehat{\omega}}^{-1}(E)$
(possibly except for dimension $d>1$ and a set of finite Liouville measure,
for which ${\overline{v}}^\pm(\widehat{\omega},x_0)=0$).
\end{remark}
In case of non--existence of the limit, we will set ${\overline{v}}^\pm(\omega,x_0):=0$.
However, typically the above limit exists.
This is shown below by invoking Birkhoff's ergodic theorem.
As this deals with finite measures, we have to change our measure space.
Namely, we set
\[\hat{P}:=P/\Theta\text.\]
\nomenclature[APhat]{$\hat P$}{quotient of $P$ by ${\mathcal L}$}{}{}%
As $P=\bigcup_{\ell\in{\mathcal L}}\Theta_\ell(\Omega\times\bR^d_p\times\cD)$
with \emph{fundamental domain} of~${\mathcal L}$
\[\cD\,:=\,\bigl\{{\textstyle \sum_{k=1}^d} x_k\ell_k\ \bigm| \ x_k\in[0,1]\bigr\}
\,\subseteq\,\bR^d_q\text,\]
the topological space~$\hat{P}$ is homeomorphic to
\begin{equation}
(\Omega\times\bR^d\times \cD)/{\sim}\text,
\Leq{homeo}
$\sim$ identifying points in $\Omega\times\bR^d_p\times\pa\cD$
via the $\Theta$ action.
Using \eqref{A} and \eqref{B}, the covering projection
\[\hat{\pi}\colon P\to\hat{P}\]
\nomenclature[Gpihat]{$\hat\pi$}{covering projection $P\to\hat P$}{}{}%
allows us to induce measures ${\hat{\mu}}$%
\nomenclature[Gmuhat]{${\hat{\mu}}$}{invariant measure on $\hat P$}{}{}
and ${\hat{\mu}}_E$%
\nomenclature[GmuhatE]{${\hat{\mu}}_E$}{invariant measure on $\hat P$}{}{}
on $\hat{P}$ by setting
\begin{equation}
{\hat{\mu}}({\hat{B}}):=\mu\bigl(\hat{\pi}^{-1}({\hat{B}})\cap(\Omega\times\bR^d_p\times\cD)\bigr)
\qquad\bigl({\hat{B}}\in\cB({\hat{P}})\bigr)
\Leq{mu:hat}
and similarly for ${\hat{\mu}}_E$ (note that the Lebesgue measure
$\lambda^d(\pa\cD)=0$).
Furthermore \eqref{B} allows us to define the continuous energy function
${\hat{H}}\colon{\hat{P}}\to\bR$%
\nomenclature[AHhat]{${\hat{H}}$}{hamiltion function on $\hat P$}{}{}
and flow ${\hat{\Phi}}\colon\bR\times{\hat{P}}\to{\hat{P}}$%
\nomenclature[GPhitilde]{${\hat{\Phi}}$}{hamiltion flow on $\hat P$}{}{}
uniquely by
$H={\hat{H}}\circ\hat{\pi}$ and ${\hat{\Phi}}^t\circ\hat{\pi}=\hat{\pi}\circ\Phi^t$.
Finally, by $\Theta$--invariance of the momenta $p\colon P\to\bR^d$,
they descend to
\begin{equation}
{\hat{p}}\colon{\hat{P}}\to\bR^d \qtextq{,} {\hat{p}}\circ\hat{\pi}=p.
\Leq{hat:p}
\nomenclature[Aphat]{${\hat{p}}$}{momentum on $\hat P$}{}{}%
\begin{proposition}\label{prop:A}
The asymptotic velocities \eqref{D} exist $\mu$--a.e.\ on~$P$, and
${\overline{v}}^+={\overline{v}}^-\ \mu$--a.e.\,.
Setting ${\overline{v}}(x):={\overline{v}}^\pm(x)$ in case of equality and ${\overline{v}}(x)=0$ otherwise,
\[{\overline{v}}\in L_{\mathrm{loc}}^\infty(P,\mu)\qtextq{,}
{\overline{v}}\circ\Phi^t={\overline{v}}\qtextq{and} {\overline{v}}\circ\Theta_\ell={\overline{v}}
\qquad(t\in\bR,\ell\in{\mathcal L})\text.\]
\end{proposition}
\begin{proof}
For initial conditions $x_0=(p_0,q_0)\in \bR^d_p\times \bR^d_q$ we have
\begin{align*
{\overline{v}}^\pm(\omega,x_0)&
=\lim_{T\to\pm\infty}\frac{q_\omega(T,x_0)-q_0}{T}
=\lim_{T\to\pm\infty}\frac{1}{T}\int_0^Tp_\omega(t,x_0)\,dt\\&
=\lim_{T\to\pm\infty}\frac{1}{T}\int_0^T{\hat{p}}\bigl(t,\hat{\pi}(\omega,x_0)\bigr)\,dt
\text.
\end{align*}
But as the finite measures ${\hat{\mu}}_E$ on ${\hat{P}}$ are ${\hat{\Phi}}$--invariant,
by Birkhoff's ergodic theorem the limits $\lim_{T\to\pm\infty}\frac{1}{T}
\int_0^T{\hat{p}}(t,\hat x_0)\,dt$ exist, coincide and are ${\hat{\Phi}}$--invariant
${\hat{\mu}}_E$--a.e.\,.
Moreover the limit function lies in $L^1({\hat{P}},{\hat{\mu}}_E)$ and is bounded by
$\sqrt{2(E-V_{\min})}$ in absolute value, as
$\norm{{\hat{p}}(t,\hat x_0)}^2\leq2 ({\hat{H}}(\hat x_0)-V_{\min})$.
Thus it is in $L^\infty({\hat{P}},{\hat{\mu}}_E)$.
The measure ${\hat{\mu}}$ is non--finite, but
${\hat{\mu}}{\upharpoonright}_{{\hat{H}}^{-1}((-\infty,E])} ={\hat{\mu}}_E$.
Thus we obtain the result.
\end{proof}
\begin{corollary}\label{cor:A}
For $\beta$--a.e.\ $\omega\in\Omega$ the asymptotic velocities
${\overline{v}}_\omega^\pm\colon \bR^{d}_p\times\bR^{d}_q\to\bR^d$ exist and are equal
$\lambda^{2d}$--a.e.\,.
\end{corollary}
\begin{proof}
This follows from \cref{prop:A} and Fubini's theorem, applied to
the measure $\mu=\beta\otimes\lambda^{2d}$ on extended phase space~$P$.
\end{proof}
Similarly, for all regular energies $E$ (see \cref{sec5} below),
the asymptotic velocities exist Liouville--almost everywhere on $H^{-1}_\omega(E)$.
\section{Distribution of Asymptotic Velocities}\label{sec3}
Next we consider the joint distribution of energy and asymptotic velocity,
using the measurable maps
\[\Gamma_\omega:=(H_\omega,{\overline{v}}_\omega)\colon\bR^{d}_p\times\bR^{d}_q\longrightarrow\bR\times\bR^d \qquad
(\omega\in\Omega)\text.\]
\nomenclature[GGammaomega]{$\Gamma_\omega$}{energy velocity map}{}{}%
We thus consider phase space regions
$\bR^d_p\times Q_n\subseteq\bR^{d}_p\times\bR^{d}_q$ with
\begin{equation*}
Q_n:=\bigl\{{\textstyle \sum_{k=1}^d} x_k\ell_k\ \bigm|\
\forall k\in\{1,\dotsc,d\}:\; -n\leq x_k<n\bigr\}
\qquad(n\in\bN)\text,
\end{equation*}
\nomenclature[AQn]{$Q_n$}{rectangle in $\bR_q^d$}{}{}%
and the normalised restrictions
\[\lambda_n\, :=\,
\frac{1}{\lambda^d(Q_n)}\lambda^{2d}{\upharpoonright}_{\bR^d_p\times Q_n}
\,=\, (2n)^{-2d}\; \lambda^{2d}{\upharpoonright}_{\bR^d_p\times Q_n}\]
\nomenclature[Glambdan]{$\lambda_n$}{normalized Lebesgue measure on~$Q_n$}{}{}%
of Lebesgue measure to these regions.
\begin{proposition}\label{prop:B}
For $\beta$--a.e.\ $\omega\in\Omega$ the {\bf energy velocity distribution}
\[\nu_\omega:=\lim_{n\to\infty}\Gamma_\omega(\lambda_n)\]
\nomenclature[Gnuomega]{$\nu_\omega$}{energy velocity distribution}{}{}%
exists in the sense of vague convergence and is independent of $\omega$.
\end{proposition}
\begin{proof}
In view of Riesz representation theorem
we have to check that for every function
$f\in C_c(\bR^{d+1})=\{g\in C(\bR^{d+1})\mid\mathrm{supp}(g)\text{ is compact}\}$
\[\lim_{n\to\infty}\int_{\bR^{d+1}}f\,d\nu_{n,\omega}\qquad \text{, with $\nu_{n,\omega}:=\Gamma_\omega(\lambda_n)$,}
\]
exists and is independent of~$\omega$.
By compactness of support of~$f$, there exists an $E_0\in\bR$ (depending
on~$f$) with $f(E,{\overline{v}})=0$ for all ${\overline{v}}\in\bR^d$ and $E\ge E_0$.
On the other hand we know that
$\norm{p}\leq\sqrt{2(E_0-V_{\min})}$ if $(p,q)\in H^{-1}([V_{\min},E_0])$.
Thus we have the estimate
\[\int_{\bR^{d+1}}\abs f\,d\nu_{n,\omega}\leq\frac{(2(E_0-V_{\min})\,\pi)
^{d/2}}{\Gamma\l(\frac{d}{2}+1\ri)}\cdot\norm[\infty]f\text,\]
\nomenclature[GGamma]{$\Gamma$}{Gamma function}{}{}%
which is uniform in $n\in\bN$ and $\omega\in\Omega$.
The function
\[g\colon\Omega\to\bR \qtextq{,} g(\omega):=\int_{\bR^{d+1}}f\,d\nu_{1,\omega}\]
is thus in $L^\infty(\Omega,\beta)$.
Moreover
\[\int_{\bR^{d+1}}f\,d\nu_{n,\omega}=\frac{1}{\setsize{{\mathcal L}_n}}\sum_{\ell\in{\mathcal L}_n}
g\bigl(\vartheta_\ell(\omega)\bigr)\]
with ${\mathcal L}_n:=\bigl\{\sum_{k=1}^dn_k\ell_k\mid n_k\in\{-n,-n+1,\ldots,n-1\}\bigr\}
\subseteq{\mathcal L}$%
\nomenclature[ALn]{${\mathcal L}_n$}{rectangle in ${\mathcal L}$}{}{}.
By assumption the probability measure~$\beta$ is ergodic w.r.t.\ the
$\vartheta$--action on~$\Omega$.
Thus a lattice version of Birkhoff's
ergodic theorem (see, \emph{e.g.}\ Keller \cite{Kel98}, Thm.\ 2.1.5) assures that
\[\lim_{n\to\infty}\int_{\bR^{d+1}}f\,d\nu_{n,\omega}\ \text{ exists\ and\
is\ $\omega$--independent}\]
$\beta$--almost surely.
\end{proof}
\vspace{2mm}
The energy--velocity distribution is the unique non--random limit measure
\[\hat{\nu}\colon\cB(\bR^{d+1})\to[0,\infty]\text.\]
\nomenclature[Gnuhat]{$\hat\nu$}{non--random energy velocity distribution}{}{}%
This measure always has a certain symmetry property:
\begin{proposition}
The energy--velocity distribution $\hat{\nu}$ is invariant w.r.t.\ the
inversion of velocity
\[I\colon\bR^{d+1}\to\bR^{d+1} \qtextq{,} (E,{\overline{v}})\mapsto(E,-{\overline{v}})\text.\]
\end{proposition}
\begin{proof}
Consider $\omega\in\Omega$ with $\nu_\omega=\hat{\nu}$
and ${\overline{v}}_\omega^+(p_0,q_0)={\overline{v}}_\omega^-(p_0,q_0)$
for $\lambda^{2d}$--a.e.\ phase space point $(p_0,q_0)\in\bR^{2d}$.
(By \cref{prop:B,cor:A} $\beta$--a.e.~$\omega$ meets these conditions.)
By \emph{reversibility} of the flow
$\Phi_\omega^t=(p_\omega^t,q_\omega^t)$, that is
\[p_\omega^t(-p_0,q_0)=-p_\omega^{-t}(p_0,q_0) \qtextq{,}
q_\omega^t(-p_0,q_0)=q^{-t}_\omega(p_0,q_0)\,\text,\]
we have ${\overline{v}}_\omega^+(-p_0,q_0)=-{\overline{v}}_\omega^-(p_0,q_0)$.
Together this gives
\[{\overline{v}}_\omega(-p_0,q_0)=-{\overline{v}}_\omega(p_0,q_0)
\qquad (\lambda^{2d}\text{--a.e.})\text.\]
On the other hand the phase space region $\bR^d\times Q_n$ as well as the
measure $\lambda_n$ on it are invariant w.r.t.\ the antisymplectic
transformation $(p,q)\mapsto(-p,q)$ on phase space.
Thus the image measures $\nu_{n,\omega}= \Gamma_\omega(\lambda_n)$
are $I$--invariant.
This carries over to the vague limit $\nu_\omega=\hat{\nu}$.
\end{proof}
\begin{defi}
For $\omega\in\Omega$ a phase space point $x_0=(p_0,q_0)\in\bR^{2d}$ is called
\emph{forward} resp.\ \emph{backward bounded} if
\[q_\omega\bigl([0,\infty),x_0\bigr) \qtextq{resp.}
q_\omega\bigl((-\infty,0],x_0\bigr)\]
are bounded subsets of configuration space $\bR^d$.
\end{defi}
Note that $\lambda^{2d}$--a.e.\ $x_0\in\bR^{2d}$ is simultaneously bounded
or unbounded in both time directions.
This is a direct consequence of the flow invariance of $\lambda^{2d}$.
\begin{proposition}
For $E>V_{\ess\max}$ and for $\beta$--a.e.\ $\omega\in\Omega$ for
every initial position $q_0\in\bR^d$ there exists an initial direction
$p_0\in\bR^d$ with $H_\omega(p_0,q_0)=E$ and $(p_0,q_0)$ forward
unbounded, with positive minimal speed (\/$\inf_{t>0}\norm{(q(t)-q_0)/t}>0$).
\end{proposition}
\begin{proof}
For $\beta$-a.a.\ $\omega\in\Omega$ we have
$V_\omega(q)\leq V_{\ess\max}$ for all $q\in\bR^d$
(see \cref{rem:A} and use the continuity of $V_\omega$).
Thus the Jacobi metric $g_{E,\omega}$%
\nomenclature[AgEomega]{$g_{E,\omega}$}{Jacobi metric}{}{}
on configuration space~$\bR^d$, given by
\begin{equation}\label{eq:JacobiMetric}
g_{E,\omega}(q)=\bigl(E-V_\omega(q)\bigr)g_\Euclid\text,
\end{equation}
is non--degenerate,
and the riemannian manifold $(\bR^d,g_{E,\omega})$ is geodesically complete,
since the conformal factor is bounded below by $E-V_{\ess\max}>0$.
So the Hopf--Rinow theorem (see, \emph{e.g.}, \cite{GHL48}, Thm.\ 2.103) implies
for any $q_n\in\bR^d$ the existence of an initial direction
$v_n\in S_{q_0}:=\{v\in\bR^d\mid g_{E,\omega}(q_0)(v,v)=1\}$%
\nomenclature[ASqzero]{$S_{q_0}$}{$g_{E\omega}$--sphere at $q_0$}{}{}
such that the forward geodesic $t\mapsto\gamma(t,q_0,v_n)$%
\nomenclature[Ggamma]{$\gamma$}{geodesic}{}{}
with initial condition $(q_0,v_n)$ meets~$q_n$ first at a time~$t_n\ge0$,
and is a shortest such geodesic so that
\[\norm{\gamma(t,q_0,v_n)-q_0}\ge \sqrt{2(E-V_{\ess\max})}\,t
\qquad \bigl(n\in\bN,\ t\in[0,t_n]\bigr)\text.\]
For $\lim_{n\ar\infty} \norm{q_n-q_0}=\infty$,
$\lim_{n\ar\infty} t_n=+\infty$, and
by compactness of $S_{q_0}$ there is an accumulation point
$v_\infty\in S_{q_0}$ of the $v_n$
leading to forward unbounded geodesic motion.
\par
Up to parametrisation, the geodesic of
$g_{E,\omega}$ with initial condition $(q_0,v_\infty)$
coincides with the trajectory $q_\omega(t,x_0)$ (with
$x_0 := \bigl(2\bigl(E-V_\omega(q_0)\bigr)\,v_\infty,q_0\bigr)$
as initial condition), see \cite[Thm.\ 3.7.7]{AM78}.
The positivity of the minimal speed is preserved.
\end{proof}
In dimensions $d\ge2$ bounded and unbounded motion can coexist
$\beta$--a.s.\ for energies $E>V_{\ess\max}$ as well as for $E<V_{\ess\max}$.
In one dimension this is not possible:
\begin{proposition}\label{prop:C}
For $d=1$ and $\beta$--a.e.\ $\omega\in\Omega$ the motion
through $x_0=(p_0,q_0)$ is
bounded if $E:=H_\omega(x_0)<V_{\ess\max}$ and
unbounded with asymptotic velocity
\begin{equation*}\label{E}
{\overline{v}}_\omega(x_0)
=\frac{\ell_1}{\bE_\beta\bigl(\tau(x_0)\bigr)}\qtextq{,}
\tau_\omega(x_0)
=\int_0^{\ell_1}\frac{\sign(p_0)}{\sqrt{2(E-V_\omega(q))}}\,dq
\end{equation*}
\nomenclature[Gtauomega]{$\tau_\omega$}{passage time for fundamental interval}{}{}%
if $E>V_{\ess\max}$.
\end{proposition}
\begin{remark}
Here the asymptotic velocity depends on $(\omega,x_0)\in P$
only via $H_\omega(x_0)$ and $\sign(p_0)$.
\Cref{abb:periodVel2}
shows the shape of $\mathrm{supp}(\hat\nu)\subseteq\bR^2$.
\end{remark}
\begin{figure}[h
\begin{center}
\ifpdf
\includegraphics[viewport=90 466 330 729,width=3cm,clip]{periodVel2.pdf}
\else
\includegraphics[viewport=90 466 330 729,width=3cm,clip]{periodVel2.ps}
\fi
\end{center}
\caption{Shape of $\mathrm{supp}(\hat\nu)\subseteq\bR^2$}
\label{abb:periodVel2}
\end{figure}
\begin{proof}
\begin{itemize}[$\bullet$]
\item For $E<V_{\ess\max}$ the connected component of $x_0$ in
$H_\omega^{-1}(E)$ is compact $\beta$--a.s.,
since there exist $q^\pm\in\bR$ with $V_\omega(q^\pm)>E$ and $q^-<q_0<q^+$.
%
\item For $E>V_{\ess\max}$ the asymptotic velocity,
if it exists for $x_0$, equals
\[{\overline{v}}(\omega,x_0)=\lim_{T\to\infty} \frac{1}{T} \int_0^T p_\omega(t,x_0)\,dt
=\lim_{n\to\infty}\frac{\int_0^{T_n(\omega)}p_\omega(t,x_0)\,dt}
{T_n(\omega)}\]
with $T_n(\omega)$ for $n\in\bN$
determined uniquely by $q_\omega(T_n(\omega),x_0)-q_0 =n\ell_1$.
Here we assume w.l.o.g.\ $\ell_1>0$ and $p_0>0$.
The numerator equals
\[\int_0^{T_n(\omega)}\dot{q}_\omega(t,x_0)\,dt = q_\omega(T_n(\omega),x_0)-q_0
=n\ell_1\text,\]
whereas
\[T_n(\omega)=\sum_{k=0}^{n-1}\int_{q_0+k\ell_1}^{q_0+(k+1)\ell_1}\frac
{1}{\sqrt{2(E-V_\omega(q))}}\,dq\]
for the denominator.
Setting
\[g\colon\Omega\to\bR \qtextq{,}
g(\omega) := \int_{q_0}^{q_0+\ell_1} \frac{1} {\sqrt{2(E-V_\omega(q))}}\,dq\text,\]
we get $T_n(\omega)=\sum_{k=0}^{n-1}$ $g\bigl(\vartheta_{k\ell_1}(\omega)\bigr)$.
By continuity of~$g$, we can
apply Birkhoff's theorem and get by the ergodicity assumption on~$\beta$
\[\lim_{n\to\infty}\frac{1}{n}T_n(\omega)=\lim_{n\to\infty}\frac{1}{n}
\sum_{k=0}^{n-1}g\bigl(\vartheta_{k\ell_1}(\omega)\bigr)=\bE(g)
\qquad \text{$\beta$--a.s.}\text.
\]
This proves the assertion for $\lambda^2$--a.e.\ $x_0\in \bR^2$
with $H_\omega(x_0) > V_{\ess\max}$.
The expression $\frac{1}{T}\int_0^Tp_\omega(t,p_0,q_0)\,dt$
is monotonically increasing in $p_0$, and our formula for
${\overline{v}}_\omega(x_0)$
is continuous in~$p_0$.
Thus it must be valid for \emph{all}~$x_0$.
\end{itemize}
\end{proof}
For $d\ge2$ it is an interesting question whether for large energies the
asymptotic velocity distribution given by~$\hat\nu$ is zero.
As the example below shows,
this is not always the case for non--trivial random potentials.
\begin{example}[Random Potential With Non-Zero Asymptotic Velocity]\label{ex:la}
We use the cutoff function
$F\in C_c^\infty(\bR^d,[0,1])$ from \cref{rem:never}
for the lattice ${\mathcal L}:=\bZ^d$.
Given single site potentials $\tW_j$%
\nomenclature[AWhatj]{$\tW_j$}{spatially restricted single site potential}{}{}
with $\mathrm{supp}(\tW_j)\subseteq\l[\frac{1}{4},\frac{3}{4}\ri]^d$, we set
$W_j(q):=\tW_j(q)-\sum_{k=1}^d\cos(2\pi q_k)F(q)$.
Then the random potential equals
\[V_\omega(q)\; = \; {\textstyle \sum_{\ell\in{\mathcal L}}}W_{\omega(\ell)}(q-\ell)
\; = \; \tV_\omega(q)-{\textstyle \sum_{k=1}^d}\cos(2\pi q_k)\]
with $\tV_\omega(q):=\sum_{\ell\in{\mathcal L}} \tW_{\omega(\ell)}(q-\ell)$%
\nomenclature[AVomegahat]{$\tV_\omega$}{spatially restricted random potential}{}{}.
Thus the Hamiltonian function $H_\omega\colon\bR^d\times\bR^d\to\bR$
takes the form
\[H_\omega(p,q) = \tV_\omega(q)+ {\textstyle \sum_{k=1}^d} H_\omega^{(k)}(p_k,q_k)\]
with $H_\omega^{(k)}\colon\bR^d\to\bR,\ (p_k,q_k)\mapsto\frac{1}{2}p_k^2-\cos
(2\pi q_k)$%
\nomenclature[AHomegak]{$H_\omega^{(k)}$}{separated hamiltonian function}{}{}.
The phase space regions
\[P_k:=\bigl\{(p,q)\in\bR^d\times(\bR^d\setminus S)\mid
\forall\,m\in\{1,\ldots,d\}\setminus\{k\}:H_\omega^{(m)}(p_m,q_m)<0\bigl\}\]
\nomenclature[APk]{$P_k$}{phase space region}{}{}%
with $S:
{\mathcal L}+\l[\ev,\dv\ri]^d$ %
\nomenclature[AS]{$S$}{enlarged lattice}{}{}%
are invariant w.r.t.\ $\Phi_\omega^t$, since $\tV_\omega(q)=0$ for
$q\in\bR^d\setminus S$, and thus the motion separates.
For all $x=(p,q)\in P_k$ the asymptotic velocity exists.
If in addition $E:=H_\omega^{(k)}(p_k,q_k)>1$,
then ${\overline{v}}^\pm(x)=(0,\ldots,0,{\overline{v}}^{(k)},0,\dotsc,0)$ with
${\overline{v}}^{(k)}:=\frac{\sign(p_k)\pi\sqrt{E-1}}{\sqrt{2}{\cal K}(2/(1-E))}$, ${\cal K}$%
\nomenclature[AK]{${\cal K}$}{complete elliptic integral of the first kind}{}{}
being the complete elliptic integral of the first kind (see \cite{AK98}).
As $E\nearrow\infty$, the intersections $P_k\cap H_\omega^{-1}(E)$ have
density w.r.t.\ Liouville measure on $H_\omega^{-1}(E)$ scaling like
$E^{-(d-1)/2}$.
Thus for no total energy strictly above $2-d$
the distribution of asymptotic velocity is concentrated in zero.
\end{example}
\section{Poisson Potentials}\label{sec4}
Compared to the lattice case handled above, potentials based on marked
Poisson fields have some new features like unboundedness and invariance
w.r.t.\ $\bR^d$--translations.
So we discuss them in this section.
Many properties should generalise to other ergodic random potentials (like
gaussian potentials).
Again we start with single site potentials~$W_j$ indexed by~$j\in J$,
$\setsize J<\infty$,
but we assume for simplicity that for some~$\eta\ge2$
\[W_j\in C_c^\eta(\bR^d,\bR)\text.\]
\nomenclature[AWj]{$W_j$}{compactly supported single site potential}{}{}%
We now consider the marked Poisson process with spac
\begin{align*}
\tilde{\Omega}
:=\bigl\{\omega\bigm|\omega&\text{ measure on
$\bigl(\bR^d\times J,\ \cB(\bR^d\times J)\bigr)$ with}\\&
\omega(K)\in\bN_0\text{ if $K\subseteq\bR^d\times J$ is compact}
\bigr\}\text.
\end{align*}
\nomenclature[GOmegatilde]{$\tilde\Omega$}{counting measures on $\bR^d\times J$}{}{}%
$\tilde{\Omega}$ is the space of all counting measures on $\bR^d\times J$
and carries the vague topology generated by the basis consisting of the sets
\begin{equation}
\hspace*{-2mm}{\cal N}_\omega(\psi_{0},\dotsc,\psi_k):=
\bigl\{\omega'\in\tilde{\Omega}\bigm|\forall i\in\{0,\dotsc,k\}\colon
\abs{\textstyle\int\psi_i\, d \omega'-\int\psi_i\, d \omega}<1\bigr\}
\Leq{eq:cN}
\nomenclature[ANomega]{${\cal N}_\omega$}{basis for vague topology}{}{}%
with $k\in\bN_{0}$, $\omega\in\tilde{\Omega}$ and $\psi_i\in C_c(\bR^d\times J,\bR)$
($i\in\{0,\dotsc,k\}$), see, \emph{e.g.}\ \cite{Rue87}.
For $j\in J$ and compact $K\in\cB(\bR^d)$ the random variables
\[N_{K,j}\colon\tilde{\Omega}\to\bN_{0}\qtextq{,}
\omega\mapsto\omega(K\times \{j\})\text,\]
\nomenclature[ANKj]{$N_{K,j}$}{particle number function}{}{}%
are called \emph{particle number functions}.
We fix intensities $\rho_j{}\ge0,\ j\in J$%
\nomenclature[Grhoj]{$\rho_j$}{intensity of Poisson process}{}{}.
Then $\beta$%
\nomenclature[Gbeta]{$\beta$}{Poisson measure}{}{}
is the unique probability measure on
$\bigl(\tilde{\Omega},\cB(\tilde{\Omega})\bigr)$ with
\begin{equation}
\beta\bigl(\{\omega\in\tilde{\Omega}\mid N_{K,j}(\omega)=m\}\bigr)
=\frac{\bigl(\rho_j\lambda^d(K)\bigr)^m}{m!\exp\bigl(\rho_j\lambda^d(K)\bigr)}
\Leq{beta:def}
for all $m\in\bN_{0}$,
$j\in J$ and $K\in\cB(\bR^d)$ with $\lambda^d(K)<\infty$.
$\bigl(\bR^d\times J,\tilde{\Omega},\cB(\tilde{\Omega}),\beta\bigr)$ is called
\emph{marked Poisson process on $\bR^d$
with marks in~$J$ and intensities~$\rho_j$}, see \cite[Chap.~4.2]{SKM87}.
It induces the random potential
\[V\colon\tilde{M}:=\tilde{\Omega}\times\bR^d \longrightarrow\bR\qtextq{,}
(\omega,q) \longmapsto \int_{\bR^d\times J} W_{j}(q-x)\, d\omega(x,j)\text.\]
\nomenclature[AV]{$V$}{Poisson potential}{}{}%
\begin{proposition}
The potential $V\colon\tilde{\Omega}\to\bR$ is continuous, and
\[V_\omega\in C^\eta(\bR^d,\bR)\qquad (\omega\in\tilde{\Omega})\text.\]
There is a $\beta$--measure zero subset $N\subseteq\tilde{\Omega}$%
\nomenclature[AN]{$N$}{exceptional set of poisson configurations}{}{}
such that for $\Omega:=\tilde{\Omega}\setminus N$ and extended phase space
$P:=\Omega\times\bR^d\times\bR^d$ the restriction
$H:=\tilde{H}{\upharpoonright}_P$ of the Hamiltonian function
\[\tilde{H}\colon\tilde{P}\to\bR \qtextq{,}
(\omega,p,q)\mapsto\eh\norm p^2+V(\omega,q)\]
induces a continuous Hamiltonian flow
\[\Phi\colon\bR\times P\to P\text.\]
\end{proposition}
\begin{proof}
\begin{itemize}[$\bullet$]
\item To show continuity of~$V$, we construct for $\vep>0$ a neighbourhood
\begin{equation*}
U(\omega,q)
\,:=\,{\cal N}_\omega(\psi_1,\ldots,\psi_k)\times B_{\delta}(q)
\,\subseteq\,\tilde M
\,=\,\tilde\Omega\times\bR^d
\end{equation*}
\nomenclature[ABdelta]{$B_\delta$}{ball of radius~$\delta$}{}{}%
of $(\omega,q){}\in\tilde\Omega\times\bR^d$, such that $\abs{V(\omega',q')-V(\omega,q)}<\vep$
for all $(\omega',q')\in U(\omega,q)$.
For radius $R:=\sup\{\norm x\mid x\in\Union_{j\in J}\mathrm{supp}(W_j)\}+1$
only the finitely many Poisson points in
\begin{equation*}
{\cS_{R,q}}:=\mathrm{supp}(\omega)\cap(B_R(q)\times J)
\equiv\{(q_1,j_1),\ldots,(q_k,j_k)\}
\end{equation*}
\nomenclature[AS]{$\cS_{R,q}$}{localised support of poisson configuration}{}{}%
can contribute to $V(\omega,q)$, and their minimal distance
\begin{equation*}
d_{\min}:=\inf \{\norm{q_m-q_n}\mid 1\leq m,n\leq k,q_m\ne q_n\}
\end{equation*}
is positive.
We will need the loci~$\cS_{R,q}^j(\omega)$ of the support of~$\omega$
with index~$j\in J$, implicitely given by
\begin{equation*}
{\cS_{R,q}} = \Union_{j\in J}\cS_{R,q}^j(\omega)\times\{j\}\text,
\end{equation*}
and a common Lipschitz constant~$L$ for the single site potentials
$(W_j)_{j\in J}$.
With
$\delta:=\min\bigl\{\frac{\vep}{2L\,\omega(\cS_{R,q})},
\frac{d_{\min}}3,1\bigr\}>0$
we choose $A_j:=B_R(q)\setminus B_\delta(\cS_{R,q}^j(\omega))$ and,
employing the Kronecker delta $\delta_{\cdot,\cdot}$
and the positive part $(\cdot)_+:=\max\{0,\cdot\}$,
functions $\psi_i\in C_c(\bR^d\times J,\bR)$ with
\begin{equation*}
\psi_i(x,j):=\begin{cases}
\bigl(1-\mathrm{dist\,}(x,A_j)/\delta\bigr)_+
&(i=0)\\
\bigl(1-\delta_{j,j_i}\mathrm{dist\,}(x,B_\delta(q_i))/\delta\bigr)_+
&(i\in\{1,\ldots,k\})
\end{cases}
\end{equation*}
Thus $\int_{{\bR^d}}\psi_i\,d\omega=(1-\delta_{0,l})\omega\{(q_i,j_i)\}$,
and by definition~\eqref{eq:cN} of ${\cal N}_\omega(\psi_0,\ldots,\psi_k)$,
all $(\omega',q')\in U(\omega,q)$ fulfil the relations
\begin{equation*}
\norm{q'-q}\leq\delta\qtextq,\omega'(A_j\times\{j\})=0
\qtextq{and}
\omega'(B_\delta(q_i)\times\{j_i\})=\omega\{(q_i,j_i)\}\text,
\end{equation*}
$j\in J$.
With this preparation we deduce
\begin{align*}&
\abs{V(\omega',q')-V(\omega,q)}\\&
\le\sum_{j\in J}\Bigabs{
\Bigl(\sum_{\tilde q\in\cS_{R,q}^j(\omega')}
W_j(q'-\tilde q)\omega'(\tilde q,j)\Bigr)
-\sum_{\hat q\in\cS_{R,q}^j(\omega)}W_j(q-\hat q)\omega(\hat q,j)
}\\&
\le\sum_{j\in J}\sum_{\hat q\in\cS_{R,q}^j(\omega)}
\Bigabs{\Bigl(\sum_{\tilde q\in\cS_{\delta,\hat q}^j(\omega')}
W_j(q'-\tilde q)\omega'(\tilde q,j)\Bigr)
-W_j(q-\hat q)\omega(\hat q,j)
}\\&
\le\sum_{j\in J}\sum_{\hat q\in\cS_{R,q}^j(\omega)}
\sum_{\tilde q\in\cS_{\delta,\hat q}^j(\omega')}
\abs{W_j(q'-\tilde q)-W_j(q-\hat q)}\omega'(\tilde q,j)\\&
\le\sum_{j\in J}\sum_{\hat q\in\cS_{R,q}^j(\omega)}
2L\delta\,\omega(\hat q,j
=2L\delta\,\omega(\cS_{R,q})
\le\ve\text,
\end{align*}
since $\norm{q'-\tilde q-(q-\hat q)}
\le\norm{q'-q}+\norm{\tilde q-\hat q}
\le2\delta$.
\item This implies that $\tilde{H}\colon\tilde{P}\to\bR$, too, is continuous
on $\tilde{P}:=\tilde{\Omega}\times\bR^{2d}$ and for all $\omega\in\tilde
{\Omega}$ the Hamiltonian $\tilde{H}_\omega\in C^\eta(\bR^{2d},\bR)$.
However, this only guarantees \emph{local} unique existence of the flow.
We now set
\begin{equation}
N:=\bigl\{\omega\in\tilde\Omega\bigm|\liminf_{\norm q\to\infty}V_\omega(q)+
c\LA q\RA^2=-\infty\ \text{for\ all}\ c>0\bigr\}
\Leq{N}
For $\omega\in\tilde\Omega\setminus N$ and $E\in\bR$ there exists a $C>0$ with
\[\sqrt{2(E-V_\omega(q))_+}\leq C\LA q\RA \qquad (q\in\bR^d)\text.\]
Thus the solution of the initial value problem with energy~$E$
exists for all times.
Namely as
\[\LA q_\omega(t)\RA\frac{d}{dt}\LA q_\omega(t)\RA\leq\norm{q_\omega(t)}\,
\norm{p_\omega(t)}\leq C\LA q_\omega(t)\RA^2\text,\]
$t\mapsto\norm{q_\omega(t)}$ is at most of exponential growth.
\item
We show that $\beta(N)=0$.
We set $W_{\max}:=\max\{\abs{W_j(q)}\mid(q,j)\in\bR^d\times J\}>0$,
${\rm diam}:=\max_j{\rm diam\,}({\rm supp\, }W_j)$
\nomenclature[AWmax]{$W_{\max}$}{supremum of single site potentials}{}{}
and $N_{B_r(q),J}:=\sum_{j\in J}N_{B_r(q),j}$.
$\beta(N)=0$ follows from $\lim_{R\to\infty}\beta(N_R)=0$ with
\begin{equation}
N_R:=\bigl\{\omega\in\tilde\Omega\bigm|\min_{\norm q\le R}V_\omega(q)\le
-R^
\bigr\}\text.
\Leq{NR}
Now for $V_\omega(q)\le -R^2$ to occur for any point
$q\in B_{\rm diam}(Q)$
we must have
\begin{equation}
N_{B_{2 {\rm diam}(Q)},J}(\omega)\ge R^2/W_{\max}.
\Leq{qQ}
As $B_R(0)$ is ${\rm diam}$--spanned (see Walters \cite{Wal82})
by a set ${\cal Q}\subseteq B_R(0)$
of points $Q$ with cardinality $\setsize{{\cal Q}}=\cO(R^d)$,
\eqref{NR} follows from \eqref{qQ} and \eqref{beta:def}.
\item
Finally, continuity of the flow~$\Phi$
follows from the theorem on continuous dependence of solutions on parameters.
\end{itemize}
\end{proof}
\begin{remark}[Comparison With Quantum Mechanics]
We have
\begin{equation}
\bE_\beta\l(\abs{V_\centerdot(0)}^{\frac{d+\vep}{2}}\ri)\leq
(W_{\max})^{\frac{d+\vep}{2}}\, \bE_\beta(N_{B_R(0),J})^{\frac{d+\vep}{2}}
\Leq{est}
Estimate \eqref{est} should be compared with a criterion for essential
self-adjointness of the corresponding random Schr\"odinger operator
$-\Delta+V$ on $L^2(\bR^d)$.
This is generally true for $\vartheta$--ergodic~$V$
if $\bE_\beta\bigl(\abs{V(0)}^{\tilde{d}}\bigr)<\infty$ for
$\tilde{d}:=2\lceil\frac{d+1}{4}\rceil$, see \cite{PF92}, Theorem 5.1.
\end{remark}
Unlike in the lattice case the Poisson potential~$V$ is invariant w.r.t.\ a
faithful
$\bR^d$--action on~$P$:
For $\ell\in\bR^d$ set
\begin{equation}\label{poissonaction}
\vartheta_\ell\colon\tilde{\Omega}\to\tilde{\Omega}\qtextq{,}
\vartheta_\ell(\omega)(A):=\omega(\{(\ell+x,j)\mid(x,j)\in A\})\text.
\end{equation}
\nomenclature[Gthetaell]{$\vartheta_\ell$}{$\bR^d$--action on $\tilde\Omega$}{}{}%
Then~$N$, defined in \eqref{N}, is $\vartheta$--invariant so that we get
the $\bR^d$ action
\[\Theta\colon\bR^d\times P\to P\qtextq{,}
(\ell,\omega,p,q)\mapsto(\vartheta_\ell(\omega),p,q-\ell)\]
\nomenclature[GTheta]{$\Theta$}{$\bR^d$--action on poissonian~$P$}{}{}%
on extended phase space, with
\[H\circ\Theta_\ell=H \qtextq{,} \mu\circ\Theta_\ell=\mu \qtextq{and}
\Phi^t\circ\Theta_\ell=\Theta_\ell\circ\Phi^t\]
for all $t \in\bR^d$ and $\ell\in\bR^d$.
In order to control the existence of the asymptotic velocities, we select
an arbitrary regular lattice ${\mathcal L}\subseteq\bR^d$, \emph{e.g.}\ ${\mathcal L}:=\bZ^d$, and set
$\Theta^{\mathcal L}:=\Theta{\upharpoonright}_{{\mathcal L}\times P}$%
\nomenclature[GThetaL]{$\Theta^{\mathcal L}$}{${\mathcal L}^d$--action on poissonian~$P$}{}{}.
Then, as in \cref{sec2} above, we consider the covering projection
\[\hat{\pi}\colon P\to\hat{P}:=P/\Theta^{\mathcal L}\]
to the factor space ${\hat{P}}$.
Like in \eqref{homeo}, ${\hat{P}}$ is homeomorphic to
$(\Omega\times\bR^d\times\cD)/{\sim}$.
On ${\hat{P}}$ we consider the measures ${\hat{\mu}}$ and ${\hat{\mu}}_E\ (E\in\bR)$, defined
like in \eqref{mu:hat}.
Again, by $\Theta$--invariance similar to \eqref{hat:p} we can define momenta
on ${\hat{P}}$ by
\[{\hat{p}}\colon{\hat{P}}\to\bR^d\qtextq{,} {\hat{p}}\circ{\hat{\pi}}=p\text.\]
But unlike in \cref{sec2}, the potential $V_\omega$
is $\beta$--a.s.\ unbounded, and thus ${\hat{H}}^{-1}((-\infty,E])$
is compact only if the $W_j$ are non--negative.
Therefore we need the following lemma:
\begin{lemma}\label{lem:L}
For all $E\in\bR$ the measure ${\hat{\mu}}_E$ on ${\hat{P}}$ is finite, and
\begin{equation}\label{eq:I}
\int_{\hat{P}}\norm{{\hat{p}}}\,d{\hat{\mu}}_E<\infty.
\end{equation}
\end{lemma}
\begin{proof}
We estimate, setting $r(q):=\sqrt{2(E-V(q))_+}$,
\[ {\hat{\mu}}_E({\hat{P}})
=\bE_\beta\l(
\int_{\cD}\int_{\bR^d}\idty_{B_{r(q)}}(p)\,dp\,dq
\ri)
=\tau_d\,\lambda^d(\cD)\,
\bE_\beta \l( 2\bigl( E-V(0) \bigr)_+^{\ d/2}\ri)
<\infty
\]
with $\tau_d:=\lambda^d\bigl(B_1(0)\bigr)=\frac{\pi^{d/2}}{\Gamma(1+d/2)}$,
\nomenclature[Gtaud]{$\tau_d$}{Lebesgue volume of unit ball in $\bR^d$}{}{}%
and similar (but with exponent $(d+1)/2$) for~\eqref{eq:I}.
\end{proof}
Thus we get the analog of \cref{prop:A}:
\begin{proposition}
The asymptotic velocities
\ ${\overline{v}}^\pm(\omega,x_0)\,:=\,
\lim_{T\to\pm\infty}\frac{q_\omega(T,x_0)}{T}$
exist and are equal $\beta$--a.s.\ on~$P$.
Furthermore
\[{\overline{v}}\in L^\infty_{\mathrm{loc}}(P) \qtextq{and} {\overline{v}}\circ\Phi^t=
{\overline{v}}\circ\Theta_\ell={\overline{v}}\text.\]
\end{proposition}
\begin{proof}
We invoke Birkhoff's theorem for
\begin{equation*}
{\overline{v}}^\pm(\omega,x_0)
=\lim_{T\to\pm\infty}\frac1T\int_0^T{\hat{p}}\bigl(t,{\hat{\pi}}(\omega,x_0)\bigr)\,dt\text,
\end{equation*}
using \cref{lem:L}.
\end{proof}
\begin{remarks}
\begin{enumerate}[1.]
\item
The $\bR^d$ action~\eqref{poissonaction} is mixing on $\tilde\Omega$,
see \cite[p.~27]{PF92}.
Thereby the action of the lattice~${\mathcal L}$ by translations
$\vartheta_\ell\ (\ell\in{\mathcal L})$ on~$\tilde\Omega$ is mixing, too,
and in particular $\beta$--ergodic.
Thus, like in \cref{prop:B}, the energy--velocity distributions
$\Gamma_\omega$ for the Poisson potentials are $\beta$--a.s.\ deterministic.
\item
For $d=1$ one gets a result similar to \cref{prop:C}.
Note, however that $\mathrm{supp}\bigl(V(\beta\otimes\lambda^d)\bigr)$
equals the closure of $(V_{\ess\min},\infty)$,
if there is a single site potential $W_j$ with $W_j(q)>0$ for some $q\in\bR$.
In that case all motion is bounded $\beta$--a.s.\,.
\end{enumerate}
\end{remarks}
\section{Singular Values of the Hamiltonian}\label{sec4a}
Since Hamiltonian motion enjoys conservation of energy,
one has to decompose phase space into energy shells
in order to find ergodic motion.
This is possible for regular energy values,
so we first study the set of singular values of the Hamilton function.
Both for the lattice and the Poisson case we have the following result:
\begin{proposition}
The closure of the set of singular values of $V_\omega$ is
$\beta$--almost surely deterministic.
\end{proposition}
\begin{proof}
For $S_{k,m}:=[m 2^{-k},(m+1) 2^{-k}]$ with $k\in\bN$ and $m\in\bZ$
the set
\[\Omega_{k,m}:=\bigl\{\omega\in\Omega\bigm|\exists q \in \bR^d:
\nabla_q V(\omega,q) =0\mbox{ and }
V(\omega,q)\in S_{k,m}\bigr\} \]
is ${\mathcal L}$-invariant. Thus by $\beta$--ergodicity it is of measure zero or one.
The sets $S_k:=\bigcup_{m\in\bZ:\, \beta(\Omega_{k,m})=1} S_{k,m}$ are closed, and
$S_{k+1}\subseteq S_k$.
\[\Omega_{k}:=\{\omega\in\Omega\bigm|{\rm CVal}_\omega\subseteq S_k \}\]
is still invariant and of measure one.
The same is true for $\Omega_{\infty}:= \bigcap_{k\in\bN}\Omega_{k}$.
The potentials indexed by $\omega\in \Omega_{\infty}$ have their critical values
in the closed set $S_\infty:=\bigcap_{k\in\bN}S_k$.
It is $\beta$--almost surely the closure of the set of singular values of~$V_\omega$.
\end{proof}
\begin{example}[Denseness of Singular Values]
The set of singular values may be dense in
$[V_{\ess\min},V_{\ess\max}]$.
This is the case
$\beta$--almost surely for the random $\bZ$--lattice potential on
$\bR$ given by the single site potentials
\[W_j(q):=\textstyle{ \sum_{\ell\in\bN}}\; j
\abs J^{-\ell}\chi\bigl(q-(-1)^\ell\lfloor\ell/2\rfloor\bigr)
\quad\bigl(q\in\bR,\;j\in J:=\{0,\ldots,\setsize J-1\}\bigr)\text,\]
with $\chi\in C^\infty_c(\bR,[0,1])$, $\chi{\upharpoonright}_{\l[-\ev,\ev\ri]}=1$
and ${\rm supp} (\chi)\subseteq[-\eh,\eh]$, if $\beta:=\otimes_{\bZ} \hat{\beta}$
is the product measure of $\hat{\beta}:=\textstyle{
\sum_{j\in J}}\delta_j/\setsize J$.
Then for any $\omega\in \Omega=J^\bZ$ with a dense
$\vartheta$--orbit the set of singular values is dense in $[0,1]$\,
and this is the case for $\beta$-a.e.\ $\omega\in\Omega$.\par
Note that the fall-off of the
single site potentials is not only polynomial as assumed in
\eqref{as:sisi} but exponential, with rate
$W_j(q)=\cO\bigl(\setsize J^{-2|q|}\bigr)$.
\end{example}
The exponential decay in this example is below
the rate that ensures measure zero for the closure of the set
${\rm CVal}_\omega =V_\omega ({\rm CSet}_\omega)$%
\nomenclature[ACVal]{$\mathrm{CVal}$}{set of singular values of~$V$}{}{}%
\nomenclature[ACSet]{$\mathrm{CSet}$}{set of singular points of~$V$}{}{}
of singular values of $V_\omega$:
\begin{proposition}
Let for $d=1$ and some $\vep>0$
the single site potentials $W_j\in C^2(\bR,\bR)$
obey (for the lattice $\bZ$)
the decay estimate $W_j'(q)=\cO\bigl(\setsize J^{-(4+\vep)\abs q}\bigr)$.
Then
\[\lambda^1\bigl(\overline{{\rm CVal}_\omega}\bigr)=0
\qquad(\omega\in\Omega)\text.\]
\end{proposition}
\begin{proof}
$\bullet$
By using the lattice translation $\vartheta$,
it suffices to show that
\begin{equation}
\lambda^1\left(\overline{
{\textstyle \bigcup_{\omega\in\Omega}}
V_\omega ({\rm CSet}_\omega \cap[0,1]) }\right)=0.
\Leq{Vla0}
For this we prove that for $R\in\bN$ large
and the middle sequence $\omega^{(R)}%
\nomenclature[GomegaR]{$\omega^{(R)}$}{middle sequence of $\omega$}{}{}
\in J^{\{-R,\ldots,R\}}$
of $\omega$ we have
\begin{equation}
V_\omega \bigl({\rm CSet}_\omega \cap[0,1] \bigr)\;\subseteq\;
M_{\omega^{(R)}}\qquad (\omega\in\Omega)\text,
\Leq{M:inclusion}
for suitable closed subsets $M_{\tau} \subseteq \bR$
of measures
\begin{equation}
\lambda^1(M_{\tau})\le \setsize J^{-(2+\vep/4)R}\qquad
\bigl(\tau\in J^{\{-R,\ldots,R\}}\bigr)\text.
\Leq{M:measure}
\nomenclature[Gtau]{$\tau$}{finite sequence in~$J$}{}{}%
As this implies
$\lambda^1\Bigl(\overline{
{\textstyle \bigcup_{\omega\in\Omega}
V_\omega ({\rm CSet}_\omega \cap}[0,1])}\Bigr)\le \setsize J^{1-\vep R/4}$,
\eqref{Vla0} then follows as $R\ar\infty$. \\
$\bullet$
$c^{(2)}:=\sup\bigl\{\sum_{j\in\bZ}\abs{W''_{\omega_j}(q-j)}\bigm
\omega\in\Omega,q\in\bR\bigr\}<\infty$ is a uniform upper bound for the second derivative of
\[\tilde{V}_\tau(q) :=
{\textstyle \sum_{\ell=-R}^R} W_{\tau(\ell)}(q-\ell)\qquad
\bigl(q\in\bR,\,\tau\in J^{\{-R,\ldots,R\}}\bigr)\text.\]
Then for $\delta\le c^{(2)}$ the disjoint union
$\{x\in [0,1]\mid\abs{\tilde{V}'_\tau(x)}\le \delta\}$ of closed intervals
has at most $i_{\max}(\tau)\le c^{(2)}/\delta$ intervals $I_1(\tau),\dotsc,I_{i_{\max}(\tau)}(\tau)$
containing a point~$q$ with $\tilde{V}'_\tau(q)=0$,
since each such interval has at least length $\delta/c^{(2)}$.
\\[2mm]
We set
$M_\tau:=\bigcup_{i=1}^{i_{\max}(\tau)} B_{\delta^2}\big(\tilde{V}_\tau(I_i(\tau))\big)$%
\nomenclature[AUdelta]{$B_\delta$}{open neighbourhood of size~$\delta$}{}{}.
This union of thickened intervals has measure
\[\lambda^1(M_\tau)\le
\lambda^1\Bigl(\tilde{V}_\tau\big({\Union\nolimits}_{i=1}^{i_{\max}
(\tau)}I_i(\tau)\big)\Bigr)
+ 2\delta^2\cdot i_{\max}(\tau)\le (1+2c^{(2)})\;\delta\text.\]
So for $\delta\equiv\delta_R:= \setsize J^{-(2+\vep/2)R}$
and $R$ large, \eqref{M:measure} is satisfied.\\
$\bullet$
By the decay assumption $|W_j'(q)|\le c \setsize J^{-(4+\vep)\abs q}$
on the derivatives of the single site potentials, for
$q\in[0,1]\setminus \bigcup_{i=1}^{i_{\max}(\tau)} I_i(\tau)$ one has
\begin{equation*}
\abs{V'_\omega(q)}
\ge\abs{\tilde{V}'_\tau(q)}-c \sum_{\ell\in\bZ:|\ell|>R}\setsize J^{-(4+\vep)\abs {\ell}}
\ge
\delta_R-c\,\delta_R^2 >0
\end{equation*}
for $R$ large enough. So
${\rm CSet}_\omega \cap[0,1]\subseteq \bigcup_{i=1}^{i_{\max}(\tau)} I_i(\tau)$
in \eqref{M:inclusion}.\\
$\bullet$
The single site potentials themselves, and not only their derivatives,
decay like $W_j(q)=\cO\bigl(\setsize J^{-(4+\vep)|q|}\bigr)$.
This means that
\[\abs{V_\omega(q)- \tilde{V}_{\omega^{(R)}}(q)}\le
c \textstyle \sum_{\ell\in\bZ:|\ell|>R}\setsize J^{-(4+\vep)\abs {\ell}} \le\delta^2
\qquad\big(q\in[0,1]\big).\]
But $\delta^2$ is the parameter appearing in the definition of $M_\tau$.
So \eqref{M:inclusion} is satisfied, too.
\end{proof}
A similar statement, but with
superexponential decay rate $\cO\bigl(\setsize J^{-(c_{\mathcal L}+\vep)\|q\|^d}\bigr)$,
should be sharp for $d\ge 1$,
with a constant $c_{\mathcal L}$ depending on the lattice
${\mathcal L}\subseteq\bR^d$.
We denote the restriction of the flow $\Phi_\omega$ to the energy surface
$\Sigma_{E,\omega}:=H_\omega^{-1}(E)\subseteq\bR^{2d}$%
\nomenclature[GSigmaEomega]{$\Sigma_{E,\omega}$}{energy surface}{}{}
by $\Phi_{E,\omega}$%
\nomenclature[GPhiEomega]{$\Phi_{E,\omega}$}{flow on~$\Sigma_{E,\omega}$}{}{}.
For regular values~$E$ of $H_\omega$,\ $\Sigma_{E,\omega}$ carries a
$\Phi_{E,\omega}$--invariant measure $\lambda_{E,\omega}$%
\nomenclature[GlambdaEomega]{$\lambda_{E,\omega}$}{Liouville measure on $\Sigma_{E,\omega}$}{}{}
derived from Lebesgue measure $\lambda^{2d}$ on phase space
(see, \emph{e.g.}\ \cite{AM78}, Thm.~3.4.12).
This non--atomic \emph{Liouville measure} thus exists for all
$E>V_{\max}$, using the equality of regular values of $H_\omega$ and $V_\omega$.
For $V_\omega\in C^{d}(\bR^d,\bR)$
by Sard's theorem (see, \emph{e.g.} \cite[Thm. 3.1.3]{Hir76}) it
in fact exists for almost all energies $E>\inf V_\omega$.
\begin{remark}[Denseness of Singular Values, Poisson Case]
For (non--trivi\-al) Poisson potentials,
$\beta$--almost surely,
the set $\overline{{\rm CVal}_\omega}$ equals $\bR$, $[0,\infty)$ or
$(-\infty,0]$, depending on the signs occuring in
the union of ranges of the single site potentials.
This is true, since
\begin{enumerate}[(a)]
\item
for any real number $r$ between zero and an extreme value of $W_j$
the sum $U_a(q):=W_j(q)+W_j(q+ae_1)$%
\nomenclature[AUa]{$U_a$}{sum of two Poisson single site potentials}{}{}
of~$W_j$ and its translate by a suitable $a={}a(r)\in\bR$ has~$r$
as a singular value. For the case $\max\; W_j=W_j(q_0)>0$ we take
\[ \mathrm{MM}(a) :=
\sup_{c\in C} \min_{t\in [0,1]} U_a\bigl(c(t)\bigr)\text,\]
\nomenclature[AMM]{$\mathrm{MM}$}{min--max function}{}{}%
with $C :=
\bigl\{c\in C^\infty\bigl([0,1],\bR^d\bigr) \bigm|c(0)=q_0, c(1)=q_0+a e_1\bigr\}$.
Then $\mathrm{MM}(0)=2\max W_j$, $\mathrm{MM}(a)=0$ for $\abs a$~large, and
$\mathrm{MM}$ is continuous.
Furthermore $\mathrm{MM}(a)$ is a singular value for~$U_a$.
The case $\min W_j=W_j(q_0)<0$ is treated by the reverse minmax problem.
\item
Integer multiples of $W_j$ are Poisson potentials, too.
\end{enumerate}
\end{remark}
\section{Notions of Ergodicity of Time Evolution}\label{sec5}
In general $\lambda_{E,\omega}$ is non--finite.
Nevertheless, we may ask whether it is an \emph{ergodic} measure (in the
sense of Aaronson \cite{Aar97}), that is,
whether the only $\Phi_\omega$--invariant measurable subsets
$A\subseteq\Sigma_{E,\omega}$ have measure zero or full measure,
i.e.\ $\lambda_{E,\omega}(\Sigma_{E,\omega}\setminus A)=0$.
Instead, we may also ask whether the restriction
\begin{equation}
{\hat{\Phi}}_E^t:={\hat{\Phi}}^t{{\upharpoonright}}_{\hat{\Sigma}_E}\ (t\in\bR)
\Leq{tPhiE}
\nomenclature[GPhihatEt]{${\hat{\Phi}}_E^t$}{flow on $\hat{\Sigma}_E$}{}{}%
of the flow ${\hat{\Phi}}$ to the (extended) energy surfaces
$\hat{\Sigma}_E:={\hat{H}}^{-1}(E)\subseteq{\hat{P}}$%
\nomenclature[GSitmahatE]{$\hat{\Sigma}_E$}{finite energy shell $\hat{\Sigma}_E$}{}{}
is ergodic.
Here the Liouville measure $\tlambda_E$%
\nomenclature[GlambdahatE]{$\tlambda_E$}{Liouville measure on~$\hat{\Sigma}_E$}{}{}
on $\hat{\Sigma}_E$, derived from the measure ${\hat{\mu}}$ on ${\hat{P}}$ by disintegration,
is finite. Ergodicity of $\hat\Phi_E$ is much cheaper obtained
than ergodicity of $\Phi_{E,\omega}$,
since on finite measure spaces Poincar\'e{'s} recurrence theorem holds,
see \cref{sect10}.
In this section we clarify the relations
between the two notions of ergodicity
and give some immediate consequences.
The results pertain to both lattice and Poisson potentials
(using quotients by $\bZ ^d$ in the latter case).
\begin{proposition}\label{prop:erg:erg}
If for energy $E\in\bR$, the flow $\Phi_{E,\omega}$ on $\Sigma_{E,\omega}$
is $\lambda_{E,\omega}$--ergodic for $\beta$--a.e.\ $\omega\in\Omega$, then
the flow ${\hat{\Phi}}_E$ on $\Sigma_E$ is ${\hat{\mu}}_E$--ergodic.
\end{proposition}
\begin{proof}
Assume that $\hat\Phi_E$ is not ergodic.
Then there exists a measurable invariant
($\tlambda_E\bigl(\hat{A}\Delta{\hat{\Phi}}_E^t(\hat{A})\bigr)=0\ \text{for\ all}\ t\in\bR$)
subset $\hat{A}\subseteq\hat{\Sigma}_E$ with
$0<\tlambda_E(\hat{A})< \tlambda_E(\hat{\Sigma}_E)$.
Its pre-image $A:={\hat{\pi}}^{-1}(\hat{A})\subseteq\Sigma_E=H^{-1}(E)$ is measurable
and invariant, too, w.r.t.\ the Liouville measure $\lambda_E$
on $\Sigma_E$.
Finally, $\lambda_E(A)>0$ and
$\lambda_E(A^c)>0$, as $A^c=\Sigma_E\setminus A=
{\hat{\pi}}^{-1}\bigl(\hat{\Sigma}_E\setminus\hat{A}\bigr)$.
\par
So $A_\omega:=A\cap\Sigma_{E,\omega}$ is $\Phi_{E,\omega}$--invariant
$\beta$--a.s., and similar for $A_\omega^c:=A^c\cap\Sigma_{E,\omega}=
\Sigma_{E,\omega}\setminus A_\omega$.
\par
By $\vartheta$--ergodicity of $\beta$ the $\beta$-a.s.\ $\vartheta$-invariant functions
\begin{equation*}
f_A,f_{A^c}\colon\Omega\to\{0,1\} \qtextq{,} f_A(\omega)=
\begin{cases}
1&\text,\ \lambda_{E,\omega}(A_\omega)>0\\
0&\text{, otherwise}
\end{cases}
\end{equation*}
(resp.\ for $A_\omega^c$) are both $\beta$--a.s.\ constant with value~$1$.
Thus $\Phi_{E,\omega}$ is not $\lambda_{E,\omega}$--ergodic $\beta$--a.s.\,.
\end{proof}
Note in passing that the flow $\Phi_E$ on $\Sigma_E$ is \emph{not}
$\lambda_E$--ergodic (given $\beta$ is not a Dirac measure on $\Omega$),
since $\Omega$ is unchanged by that flow.
Next we ask about the dynamical consequences of ergodicity.
\begin{proposition}\label{prop:P}
Given $E\in\bR$ and $\omega\in\Omega$ such that $\Phi_{E,\omega}$ is $\lambda_{E,\omega}$--ergodic
on the regular energy surface $\Sigma_{E,\omega}$,
\begin{compactitem}
\item then the asymptotic velocity ${\overline{v}}(x)=0$
for a.e.\ $x\in{}\Sigma_{E,\omega}$,
\item but the motion is unbounded a.e.\ on $\Sigma_{E,\omega}$
unless $\Sigma_{E,\omega}$ is compact.
\end{compactitem}
\end{proposition}
\begin{remark}
Note that $\Sigma_{E,\omega}$ is non--compact for $E>V_{\max}$
and $\beta$--a.s.\ non--compact for $E>V_{\ess\min}$, given ergodicity of~$\Phi_{E,\omega}$.
\end{remark}
\begin{proof}
Remember that ${\overline{v}}(x)=0$ by definition if ${\overline{v}}^\pm(x)$ do not exist or are
unequal.
\begin{itemize}[$\bullet$]
\item Assume that $\lambda_{E,\omega}(\{x\in\Sigma_{E,\omega}\mid
{\overline{v}}(x)\neq0\})>0$.
Then there exists an open half space $H_s:=\{q\in\bR^d\mid\LA q,s\RA>0\}$%
\nomenclature[AHs]{$H_s$}{half space}{}{}
indexed by a unit vector $s\in S^{d-1}$%
\nomenclature[ASd]{$S_d$}{$d$--dimensional sphere}{}{}
so that
\[\lambda_{E,\omega}(B_s)>0\qtextq{for}
B_s:=\{x\in\Sigma_{E,\omega}\mid{\overline{v}}(x)\in H_s\}\text.
\]
By reversibility of the flow and $\overline v^-=\overline v^+$
\[\lambda_{E,\omega}(B_{-s})=\lambda_{E,\omega}(B_s)\text.\]
On the other hand, ${\overline{v}}\circ\Phi^t={\overline{v}}\quad(t\in\bR)$, so that $B_s$ and
$B_{-s}$ are disjoint invariant subsets of positive measure.
Thus $\Phi_{E,\omega}$ is not $\lambda_{E,\omega}$--ergodic.
\item Assume that there exists a subset $B\subseteq \Sigma_{E,\omega}$ of
points leading to forward bounded motion with $\lambda_{E,\omega}(B)>0$.
Then for some $r>0$ the subset
\[B_r:=\{(p,q)\in B\mid\norm q\leq r\}\]
is of positive measure, too.
Furthermore, for some $r'>r$ the same applies to
\[B_{r,r'}:=\{x\in B_r\mid\norm{q_t(\omega,x)}\leq r'
\text{ for\ all}\ t\ge0\}\text.\]
Finally the set
\[{\tilde{B}}:=\bigcup_{t\ge0}\Phi_{E,\omega}^t(B_{r,r'})\subseteq \Sigma_E\]
is $\Phi_{E,\omega}$--invariant w.r.t.\ $\lambda_{E,\omega}$, and
$\lambda_{E,\omega}({\tilde{B}})>0$.
But by non--compactness of~$\Sigma_{E,\omega}$,
its complement ${\tilde{B}}^c=\Sigma_{E,\omega}\setminus{\tilde{B}}$
is of measure $\lambda_{E,\omega}({\tilde{B}}^c)>0$, since~${\tilde{B}}$
is bounded: $\norm q\leq r'$ if $(p,q)\in{\tilde{B}}$.
Thus $\Phi_{E,\omega}$ is not $\lambda_{E,\omega}$--ergodic.
\end{itemize}
\end{proof}
In one--dimensional natural mechanical systems with Hamiltonian
$H\colon\bR^2\to\bR,\ H(p,q)=\frac{1}{2}p^2+V(q)$ ergodicity on regular energy
surfaces $\Sigma_E=H^{-1}(E)$ is widespread and occurs for compact as well
as non--compact $\Sigma_E$ (examples: $V(q)=q^2,\ V(q)=q$).
In our present context, however, ergodicity in one dimension is exceptional.
\begin{proposition}
For $d=1$ and $E>V_{\ess\min}$ the flow $\Phi_{t,\omega}$ on a
regular energy surface $\Sigma_{E,\omega}$ is $\beta$--a.s.\ not
$\lambda_{E,\omega}$--ergodic.
\end{proposition}
\begin{proof}
We assume that $\omega$ is chosen so that ${\overline{v}}^+(x)={\overline{v}}^-(x)$ for
$\lambda_{E,\omega}$--a.e.\ $x\in\bR$.
This assumption is valid $\beta$--a.s.\,.
For $E<V_{\ess\max}$ the motion is bounded $\beta$--a.s., whereas for
$E>V_{\ess\max}$ asymptotic velocity is non--zero
$\beta$--a.s.\ according to \cref{prop:C}.
Both statements contradict \cref{prop:P} (note that
$\Sigma_{E,\omega}$ is non--compact $\beta$--a.s.\ for $E>V_{\ess\min}$).
But for the case left, $E=V_{\ess\max},\ \Sigma_{E,\omega}$ must be
connected, regular and non--compact.
Thus it must be diffeomorphic to $\bR$.
This can only occur if there is a $q_0$ so that $V(q)>E$ for all $q\in
(q_0,\infty)$ and $V(q)<E$ for all $q\in(-\infty,q_0)$ or vice versa.
This, however, would contradict our initial assumption ${\overline{v}}^+(x)={\overline{v}}^-(x)$.
\end{proof}
\Cref{ex:la} together with \cref{prop:P} shows that for any
dimension~$d$ there are non--trivial random lattice potentials that lead to
non--ergodic motion on the regular energy surfaces $\Sigma_{E,\omega}$ for
\emph{any}~$E$ and $\omega\in\Omega$.
The statements of \cref{prop:P} can be proven under a different
assumption, which by \cref{prop:erg:erg} is weaker than
$\lambda_{E,\omega}$--ergodicity assumption in \cref{prop:P}
\emph{for $\beta$-a.e.}~$\omega$:
\begin{proposition}\label{prop:ConsequencesOfErgodicity}
If ${\hat{\Phi}}_E$ is ${\hat{\mu}}_E$--ergodic on the regular energy surface $\hat{\Sigma}_E$,
\begin{compactitem}
\item then the asymptotic velocity ${\overline{v}}(x)=0$ for a.e.\ $x\in{}\Sigma_{E,\omega}$,
\item but the motion is unbounded a.e.\ on $\Sigma_{E,\omega}$
\end{compactitem}
for $\beta$-a.e.\ $\omega$.
\end{proposition}
\begin{proof}
Still, ${\overline{v}}(x)=0$ by definition,
where ${\overline{v}}^\pm(x)$ do not exist or are unequal.
We define $\hat v_E\colon\hat{\Sigma}_E\to\bR^d$ by
$\hat v_E\circ\hat\pi={\overline{v}}{\upharpoonright}_{\Sigma_E}$.
\begin{itemize}[$\bullet$]
\item Assume that there exists a measurable $\Omega'\subseteq\Omega$
with $\beta(\Omega')>0$ such that
$\lambda_{E,\omega}(\{x\in\Sigma_{E,\omega}\mid{\overline{v}}(x)\ne0\})>0$
for all $\omega\in\Omega'$.
This implies
\begin{equation*}
\tlambda_E(\{x\in\hat{\Sigma}_E\mid\hat v_E(x)\ne0\})>0\text.
\end{equation*}
Again, there exists an open half space $H_s:= \{q\in\bR^d\mid\LA q,s\RA>0\}$
indexed by a unit vector $s\in S^{d-1}$ so that
\begin{equation*}
\tlambda_E({\hat B}_s)>0\qtextq{for}
{\hat B}_s:=\{x\in\hat{\Sigma}_E\mid\hat v_E(x)\in H_s\}\text.
\end{equation*}
By reversibility of the flow and $\overline v^-=\overline v^+$
\[\tlambda_{E,\omega}({\hat B}_{-s})=\tlambda_{E,\omega}({\hat B}_s)\text.\]
On the other hand, $\hat v_E\circ{\hat{\Phi}}^t=\hat v_E\quad(t\in\bR)$,
so that ${\hat B}_s$ and ${\hat B}_{-s}$
are disjoint invariant subsets of positive measure.
Thus ${\hat{\Phi}}_E$ is not $\tlambda_E$--ergodic.
\item We have to show, that the bounded orbits in~$\Sigma_E$
carry no measure. Fix $r>0$ and let
\begin{equation*}
B{{}\equiv B_r}:=\{x\in\Sigma_E\mid\forall t\in\bR\colon\norm{q_t(x)}<r\}
\end{equation*}
be the set of orbits bounded by~$r$.
With help of the projections
\begin{equation*}
\pi_{\Omega,E}\colon\Sigma_E\to\Omega
\qtextq{and}
\pi_\Lambda\colon\Omega\to\Omega_\Lambda:=J^\Lambda
\end{equation*}
\nomenclature[GpiOmegaE]{$\pi_{\Omega,E}$}{projection $\Sigma_E\to\Omega$}{}{}%
\nomenclature[GpiLambda]{$\pi_\Lambda$}{projection $\Omega\to\Omega_\Lambda$}{}{}%
for finite subsets $\Lambda\subseteq{\mathcal L}$%
\nomenclature[GLambda]{$\Lambda$}{finite subset in~${\mathcal L}$}{}{}
we can partition $B=\Union_{\sigma\in\Omega_\Lambda}B_\sigma$ with
\begin{equation*}
B_\sigma:=B\isect(\pi_\Lambda\circ\pi_{\Omega,E})^{-1}\{\sigma\}
\qquad(\sigma\in\Omega_\Lambda)\text.
\end{equation*}
Since the probability measure~$\beta$ is non--atomic, so is $\lambda_E$.
We thereby conclude
\begin{equation}\label{eq:Bsigmatozero}
\sup_{\sigma\in\Omega_\Lambda}\lambda_E(B_\sigma)
\xto{\Lambda\to{\mathcal L}}0\text.
\end{equation}
On the other hand $B_\sigma$ is $\Phi_E$--invariant,
and $\hat B_\sigma:=\hat\pi(B_\sigma)$ is preserved by ${\hat{\Phi}}_E$.
By ergodicity of~${\hat{\Phi}}_{E}$ and \eqref{eq:Bsigmatozero}
we find a finite subset $\Lambda\subseteq{\mathcal L}$ such that
\begin{equation*}
\sup_{\sigma\in\Omega_\Lambda}\tlambda_E(\hat B_\sigma)=0\text.
\end{equation*}
But $\hat\pi$ is non--singular, i.e.\ preserves sets of measure~$0$,
and we conclude
\begin{equation*}
\lambda_E(B)
=\sum_{\sigma\in\Omega_\Lambda}\lambda_E(B_\sigma)
=0\text.
\end{equation*}
\end{itemize}
\end{proof}
\section{No Ergodicity in the Poisson Case}\label{sec5a}
We now assume for the Poisson case that the single site potentials are smooth,
in addition to being compactly supported.
Then ergodicity of the dynamics is atypical:
\begin{theorem}
\label{thmPoisson}
Consider a random Poisson potential on $\bR^d$.
Then for any $E\in\bR$ the motion on the energy surface
$\Sigma_{E,\omega}$ is $\beta$--a.s.\ not ergodic.
\end{theorem}
\noindent
{\bf Strategy of proof and first steps}\\
$\bullet$
The theorem is true for $d=1$ dimensions. If there is a single site
potential, say $W_1$, and $q\in\bR$ with $W_1(q)>0$, then we have
bounded orbits $\beta$--a.s.\ for all~$E$. If instead all $W_j$ are
non--positive, $\Sigma_{E,\omega}$ has two connected components for
$E>0$ and one has $\beta$--a.s.\ bounded orbits for $E\le 0$.\\
$\bullet$
So we assume $d\ge2$.
If the Poisson potential is zero, then we have free motion,
which is not ergodic in any dimension.
Otherwise there is a non--zero
single site potential, say $W_1$.
The proof method depends on the sign of
\begin{equation}
I:={\textstyle \int_{\bR^d}} W \,d\lambda^{d}
\Leq{def:I}
\nomenclature[AI]{$I$}{integral over $W_1$}{}{}%
(we temporarily omit the index of $W_1$).
Then, for given $I$ and energy~$E$, we construct a set~${\mathcal L}$%
\nomenclature[AL]{${\mathcal L}$}{set of translation points for poisson constructions}{}{}
and thereby a \emph{finite} sum $q\mapsto \sum_{\ell\in{\mathcal L}} W(q-\ell)$
of translated potentials which for the given
energy confines trajectories of positive measure.\\
This then suffices to show that $\beta$--a.s.\
the flow on $\Sigma_{E,\omega}$ is not ergodic:
\begin{enumerate}[1.]
\item
If the restriction of $\omega$
to a large ball $B_{R}(0)\subseteq\bR^d$ is near to
$\sum_{\ell\in {\mathcal L}}\delta_{(\ell,1)}$ w.r.t.\ the
topology from \eqref{eq:cN}, then the flow is shown to be
non--ergodic, too. This event
in $\Omega$ has positive probability.
\item
By the nature of the Poisson process $\beta$
the Borel--Cantelli Lemma can be applied to the
$\vartheta_\ell$--translates of that
event with $\ell\in 2R\bZ^d$,
so that the flow is even $\beta$--a.s.\ non--ergodic.
\end{enumerate}
\nomenclature[ASOd]{$\mathrm{SO}(d)$}{special orthogonal group}{}{}%
${\rm SO}(d)$--invariant potentials
$U\in C^\infty_c\bigl(\bR^d,[0,\infty)\bigr)$%
\nomenclature[AU]{$U$}{$\mathrm{SO}(d)$--invariant function}{}{}
lead to integrable motion, and for suitable~$U$ to a positive measure
of bounded orbits
(see, {\em e.g.}, Arnol'd \cite{Arn78}, Section 8). \\
We first approximate a given such
$U\in C^\infty_c\bigl(\bR^d,[0,\infty)\bigr)$,
with $U(q)=0$ for~$q$ near zero,
by summing scaled translates of the single site function
$W\in C^\infty_c(\bR^d,\bR)$.
We have $\int_{\bR^d} W_a \,d\lambda^{d}=I$ for
\nomenclature[AWa]{$W_a$}{$W$ scaled by~$a$}{}{}%
\begin{equation}\label{W-scaling}
W_a(q) := a^{-d} W(q/a)\qquad (a>0)\text.
\end{equation}%
\begin{lemma}\label{lem:dens}
For $I>0$ in \eqref{def:I} and $U$
as above there exists a map $g\in C^{\infty}(\bR^d,\bR^d)$
which, restricted to the interior of ${\rm supp}(U)$,
is a diffeomorphism onto its (bounded) image $\mathrm{Im}$%
\nomenclature[AIm]{$\mathrm{Im}$}{image of~$U$}{}{}
such that the distributions
\[D_\vep\;:= \;\vep^d \;
{\textstyle \sum_{\ell\in\vep \bZ^d \cap \mathrm{Im}}}
\;\delta_{g^{-1}(\ell)} \qquad(\vep>0)\]
\nomenclature[ADepsilon]{$D_\vep$}{distribution approximating $D_0$}{}{}%
\nomenclature[Gdeltax]{$\delta_x$}{Dirac measure on~$x$}{}{}%
converge vaguely to the distribution $D_0:=U\lambda^d$%
\nomenclature[ADzero]{$D_0$}{distribution with density~$U$}{}{}
with density~$U$ w.r.t.\ Lebesgue measure $\lambda^d$ for $\vep\searrow0$.
Moreover, in uniform $C^k$ norm, with $c:=1/(2(d+k+1))$,
\begin{equation}\label{faltung}
I^{-1}\lim_{\vep\searrow 0}D_\vep * W_{\vep^c}\; =\; U.
\end{equation}
\end{lemma}
\begin{proof}
Write $U(q)=\tilde U(\norm q)$ and set
$\tilde g(r) := \bigl(d\int_0^r x^{d-1}\tilde U(x)\,dx\bigr)^{1/d}$.
Then, by the assumptions on $U$, the map
$g\colon\bR^d\ar\bR^d$, with $g(0):=0$ and $g(q):=\tilde g(\|q\|)
\frac{q}{\norm q}$ else, is zero in a neighbourhood of zero,
$g\in C^k$ and bounded.
As $Dg(q)=\tilde g(\norm q)(\idty-P_q)/\norm q+\tilde g'(\norm q)P_q$
with the orthogonal projection~$P_q$%
\nomenclature[APq]{$P_q$}{orthogonal projection}{}{}
onto ${\rm span}(q)$%
\nomenclature[Aspan]{$\mathrm{span}$}{linear span}{}{},
$\det(Dg)(q)
=\tilde g'(\norm q)\;\bigl(\tilde g(\norm q)/\norm q\bigr)^{d-1}
=U(q)$.
This shows that $g{\upharpoonright}_{{\rm int}({\rm supp}(U))}$
is a diffeomorphism onto ${\rm Im}$.
The image measure of $\idty_{{\rm Im}}\,\lambda^d$ under $g^{-1}$
is $\det(Dg)\,\lambda^d$, and $\vep^d \;
{\textstyle \sum_{\ell\in\vep \bZ^d}}
\;\delta_{\ell} \to \lambda^d$.
So $\lim_{\vep\searrow 0} D_\vep=D_0$.
To prove the statement~\eqref{faltung} on the $C^k$~norm,
we notice that $\vep\mapsto I^{-1}W_{\vep^c}$
is an approximation to the delta distribution~$\delta_0$, and
\begin{equation}
I^{-1}{\textstyle \int_{\mathrm{Im}}}\; W_{\vep^c} \bigl(q-g^{-1}(x)\bigr)\,dx
\;=\; U(q)+\cO(\vep^c),
\Leq{U:app}
with corresponding formulae for the derivatives.
Then~\eqref{faltung} follows by a direct estimate.
\end{proof}
We need to scale natural Hamiltonian systems
\[H\colon P_{\rm ext}\ar\bR\quad , \quad H(t,p,q)=\eh\|p\|^2+V(q)\]
\nomenclature[APext]{$P_{\mathrm{ext}}$}{extended phasse space}{}{}%
with flows
on extended phase space $P_{\rm ext}:=\bR_t\times T^*\bR^d_q$.
\nomenclature[ATstar]{$T^*$}{cotangential functor}{}{}%
We use three types of scalings which all transform the natural Hamiltonian
and its flow in a simple way:
\begin{enumerate}[1.
\item \textbf{Motions}
$M:\bR^d_q\to \bR^d_q$,
\nomenclature[AM]{$M$}{motion in $\bR_q^d$}{}{}%
$q\mapsto Oq+v$ with $O\in{\rm SO}(d)$, $v\in\bR^d$
lift to symplectic maps $M^*$ on $T^*\bR^d_q$,
\nomenclature[AMstar]{$M^*$}{motion on $T^*\bR_q^d$}{}{}%
resp.\ ${\cal M}:= {\rm Id}_t\times M^*$ on $P_{\mathrm{ext}}$,
\nomenclature[AM]{${\cal M}$}{motion on $P_{\mathrm{ext}}$}{}{}%
and
\nomenclature[AMtilde]{$\tilde{\cal M}$}{motion in $C^k(P_{\mathrm{ext}},\bR)$}{}{}%
\[\tilde {\cal M}: C^k(P_{\rm ext},\bR)\to C^k(P_{\rm ext},\bR)\qtextq,
\tilde{\cal M}\, H:= H\circ {\cal M}\text.
\]
\item \textbf{Spatial scalings ($L^\infty$ dilations)}
$S_c:P_{\rm ext}\ar P_{\rm ext}$,
\nomenclature[ASc]{$S_c$}{spatial scaling}{}{}%
$S_c(t,p,q):=(t/c,\,p,\,c\,q)$, and
\nomenclature[AStildec]{$S_c$}{spatial scaling on $C^k(P_{\rm ext},\bR)$}{}{}%
\[\tilde S_c\colon C^k(P_{\rm ext},\bR)\to C^k(P_{\rm ext},\bR)\qtextq,
\tilde S_c\,H := H\circ S_c\qquad (c>0)\text.
\]
\item \textbf{Energy scalings}
$\cE_e\colon P_{\rm ext}\ar P_{\rm ext}$,
\nomenclature[AEe]{$\cE_e$}{energy scaling on $P_{\rm ext}$}{}{}%
$\cE_e(t,p,q):=\bigl(\sqrt{e}\,t,\,p/\sqrt{e},\,q\bigr)$, and
\nomenclature[AEetilde]{$\tilde\cE_e$}{energy scaling on $C^k(P_{\rm ext},\bR)$}{}{}%
\[\tilde\cE_e:C^k(P_{\rm ext},\bR)\to C^k(P_{\rm ext},\bR)\quad\text,\quad
\tilde\cE_e\, H := e\;H\circ \cE_e \qquad (e>0)\text.
\]
Notice that, regarding a single site potential~$W$
as a function on~$P_{\rm ext}$,
$W_a$ in \eqref{W-scaling}~is a combination of spatial and energy scalings of~$W$.
\end{enumerate}
\begin{proof}[The Case $I > 0$]
$\bullet$
Given $E\in\bR^+$, we choose $\tilde U\in C^k_c\bigl([0,\infty),\bR\bigr)$
with $\tilde U(r)=0$ for all~$r$ smaller than
$2\,{\rm diam}\bigl({\rm supp}(W)\bigr)$ and having one maximum,
with height $\max(\tilde U)\ge 2E$.
$U=\tilde U\circ\norm{{}\cdot{}}$ then has the property that
for all $e\in(0,E]$
the energy surface $\{(p,q)\in T^*\bR^d\mid \eh\|p\|^2+U(q)=e\}$ has
two components.\\
$\bullet$
By \cref{lem:dens} for some $\vep\in(0,\min\{1,I^{((1-c)d)^{-1}\}}]$
we find an approximation of~$U$ of the form
$I^{-1} D_\vep * W_{\vep^c}$ with the same property.
Then, by the combination of spatial and energy scalings, the sum
\begin{equation*}\textstyle
\sum_{\ell\in\vep \bZ^d \cap\mathrm{Im}}
W\bigl({}\cdot{}-\tfrac{g^{-1}(\ell)}{\vep^c}\bigr)
=\tilde\cE_{I\vep^{-(1-c)d}}\circ\tilde S_{\vep^c}
\bigl(\vep^{-d}D_\vep * W_1\bigr)
\end{equation*}%
of translated single site potentials
has that property for all energies~$e\in(0,I\vep^{-(1-c)d}E]$, too.
Since $\vep^{(1-c)d}\le I$, $E\in(0,I\vep^{-(1-c)d}E]$.\\
$\bullet$
For any realization $\omega\in\Omega$
which inside a ball of large radius
is near enough to $\sum_{\ell\in{\mathcal L}} \delta_{(\ell,1)}$, with ${\mathcal L}$ as above,
the energy surface
$\Sigma_{E,\omega}$ consists of at least two connected components.
Thus the motion on $\Sigma_{E,\omega}$ is not ergodic $\beta$--a.s..
\end{proof}
\begin{proof}[The Case $I<0$]
$\bullet$
We first treat the case $d=2$ and then indicate the modifications
needed for larger dimensions.
We use \cref{lem:dens} in order to approximate a
{\em non--positive} centrally symmetric function~$U$ by
summing scaled translates of the single site function~$W$.
We first assume that the energy is positive, more
specifically $E=1$. We write again $U(q)=\tilde U(\norm q)$ and
choose the profile $\tilde U$ so that there is a circular
periodic orbit $t\mapsto q(t)=r\V{\cos(\omega t)\\\sin(\omega t)}$
of energy~$E$ with $\omega>0$%
\nomenclature[Gomega]{$\omega$}{angular frequency}{}{}
in the potential~$U$.
\begin{figure}
\begin{center}
\ifpdf
\includegraphics{ueff.pdf}
\else
\includegraphics{ueff.ps}
\fi
\end{center}
\caption{The effective radial potential $\tilde U_\ell$}
\label{fig:radial}
\end{figure}
We thus consider the effective potential
$\tilde U_\ell(r) := \tilde U(r)+ \frac{\ell^2}{2r^2}$%
\nomenclature[AUtildel]{$\tilde U_\ell(r)$}{effective potential}{}{}
with angular momentum parameter $\ell\in\bR$%
\nomenclature[Al]{$\ell$}{angular momentum parmeter}{}{}.
The conditions for a circular periodic orbit of radius~$r$ are
$E=\tilde U_\ell(r)$ and $\tilde U'_\ell(r)=0$.
\par
In order to control the stability of the orbit
(make the linearized flow elliptic and let the frequency
vary with the perturbation), we demand $\tilde U''_\ell(r)>0$
and, say $\tilde U^{(4)}_\ell(r)\neq 0$.
All these conditions can be satisfied by, {\em e.g.},
first choosing $r:=1$, $\ell:=2$, and then finding an appropriate
$\tilde U\in C^\infty_c\bigl([0,\infty),[-1,0]\bigr)$ with $\tilde U(r)=0$
in a neighborhood of zero, $\tilde U(1)=-1$, meeting the assumptions
on the derivatives and angular frequency $\omega=2$.
Cf.~\cref{fig:radial}.\\
$\bullet$
Then by KAM theory the Hamiltonian flows with potentials $C^k$--near
to $U$ have, too, an elliptic orbit of energy $E=1$, surrounded
by invariant tori of positive measure.
See, {\em e.g.}, Arnol'd \cite{Arn78}, Appendix~8.
According to P\"oschel \cite{Po82} the
differentiability condition $W\in C^k(\bR^d,\bR)$
with $k=3d$ suffices.
Let $\vep\in(0,1)$ be small enough so that the potential
$\hat{U}:= |I|^{-1} D_\vep * W_{\vep^c}$%
\nomenclature[AUhat]{$\hat U$}{approximation to~$U$}{}{}
(with $D_\vep$ from \cref{lem:dens}) meets that condition for KAM theory.
Then with the scaling \eqref{W-scaling} of $\hat{U}$,
$|I|\vep^{-(1-c)d}\hat{U}_{\vep^{-c}}$ is a sum of
translated single site potentials meeting that condition for
energy $e:=|I|\vep^{-(1-c)d}$%
\nomenclature[Ae]{$e$}{energy}{}{}.
This proves the claim for large enough energies~$e$.\\
$\bullet$
In order to solve the problem for an arbitrary smaller energy $E<e$,
we first add to $U$ a function
$\Delta U\in C^\infty_c\bigl(\bR^d,(-\infty,0]\bigr)$,
\nomenclature[AUDelta]{$\Delta U$}{non--positive function}{}{}%
which has a constant value
$\Delta E:=e^{-1}E-1=\abs I^{-1}\vep^{(1-c)d}(E-e)<0$
on the support of~$U$.
Over $\mathrm{supp}(U)$, the flow at energy~$1$ with potential~$U$
equals the flow for energy $\abs I^{-1}\vep^{(1-c)d}E$
with potential $U+\Delta U$.\\
$\bullet$
The case of dimension $d\ge3$ cannot be treated in complete analogy,
since then motion in phase space generated by rotations
from $\mathrm{SO}(d)$ not only leaves the Hamiltonian invariant, but
also can transform a circular orbit to {\em different} circular
orbits. This shows that in the integrable problem periodic
orbits of constant radius have degenerate transverse frequencies.
KAM theory can be applied after lifting this degeneracy.
One way to do this is to construct a potential which
near the circle
\[\{q\in\bR^d\mid q_1^2+q_2^2=1, q_3=\ldots= q_d=0\}\]
is of the form $\tilde U(q_1^2+q_2^2)+ \hat U(q_3,\ldots ,q_d)$,
with
$\hat U(q_3,\ldots,q_d)=\sum_{m=3}^d(\omega_mq_m^2+\tau_mq_m^4)$%
\nomenclature[AUhat]{$\hat U$}{perturbation to lift degeneracies}{}{}
integrable, non--resonant and non--degenerate.
Since such a~$U$ locally is the sum of functions of the
coordinates, $U$~can again be locally
approximated in a manner similar to \cref{lem:dens}.
\end{proof}
\begin{proof}[The Case $I = 0$]
$\bullet$ Energies $E\le 0$ again lead to bounded motion.\\
$\bullet$ By injectivity of Radon transform (see, \emph{e.g.}, Natterer
\cite{Nat}, Theorem 2.1), there exists a hyperplane ${\cal H}\subseteq\bR^d$%
\nomenclature[AH]{${\cal H}$}{hyperplane}{}{}
(perpendicular to some $e_1\in S^{d-1}$)
so that $I':=\int_{\cal H} W \,d\lambda^{d-1}\neq 0$.
By a translation of all the single site potentials $W_j$
(which does not change the class of Poisson Hamiltonians)
we can assume that $0\in{\cal H}$.
By our assumption
$I=\int_{\bR^d} W \,d\lambda^{d}= 0$ we can even assume that
$I'>0$%
\nomenclature[AJ]{$J$}{integral of~$W$ over~${\cal H}$}{}{}.
We supplement $e_1$ to an orthonormal basis $e_1,\ldots,e_d$ of~$\bR^d$%
\nomenclature[Ae1]{$e_1,\ldots,e_d$}{Basis of $\bR^d$}{}{}
and define the lattice ${\mathcal L}:={\rm span}_\bZ(e_2,\ldots,e_d)$.
Then for $r>0$ large and the ball $B_r^d\subseteq \bR^d$ of radius~$r$
the linear combination
\begin{equation}
\widehat{W}_\vep(q):=
\vep^{d-1}\sum_{\ell\in\vep {\mathcal L}\cap B_{2r}^d} W(q-\ell)\qquad(q\in\bR^d)
\Leq{Weps}
\nomenclature[AWtildeepsilon]{$\widetilde W_\vep$}{combination of single site potentials}{}{}%
converges, as $\vep\to0$, in $C^k$ sense to
$\widehat{W}_0\in C^\infty_c(\bR^d,\bR)$%
\nomenclature[AWtildezero]{$\widetilde W_0$}{limit of $\widehat W_\vep$}{}{}
with $\widehat{W}_0(q)=\int_{{\cal H}+q_1e_1} W(x)\,dx$
for all~$q$ with $\norm{q-q_1e_1}\le r$ (setting $q_1:=\LA q, e_1\RA$).
So inside that cylinder ${\cal H}\cap B_r\,+\,{\rm span}_\bR(e_1)$,
$\widehat{W}_\vep$ is nearly invariant under translation
perpendicular to~$e_1$, and $\widehat{W}_\vep(q)>0$ if $q_1=0$.
To orbits of regular energy $E\in(0,I')$
entering ${\rm supp}\,\widehat{W}_\vep$
with velocity nearly parallel to $e_1$ and position $q$ within the
cylinder, the potential acts like nearly planar mirror.
By translating two such potentials by $\pm R e_1$ with $R\gg r$,
we get a system of two mirrors. Between these mirrors
trajectories of appropriate energy bounce back and forward
near the axis ${\rm span}_\bR(e_1)$
for a long time.
To make the motion bounded for a set of initial
conditions of positive measure, we give the mirrors
inside curvatures stricly smaller than $1/R$.
This can be done by changing the summand in \eqref{Weps} to
$W(q-\ell-Q(\ell)e_1)$ with an appropriate quadratc form $Q$%
\nomenclature[AQ]{$Q$}{quadratic form}{}{}.
Under this condition the orbit on the axes
becomes linearly elliptic for a potential with perfect
axial symmetry. Then by KAM theory and appropriate scaling we
get the result for all positive energies.
\end{proof}
\section{No Hyperbolicity for Bounded Potentials}\label{sec6}
By far the simplest ergodic flows or maps are the uniformly hyperbolic ones.
Examples include hyperbolic torus automorphisms
and geodesic flows on
compact manifolds of negative sectional curvature.
Although is known that there exist severe \emph{topological} obstructions
against a flow on a manifold to be Anosov, these do, as shown in \cref{exa}
below, not apply to the motion
in a potential on configuration space $\bR^d$.
However, by \cref{theo} below,
\emph{geometric} obstructions exist if the potential is bounded.
There exist examples of ergodic motion in smooth bounded potentials
(which are not uniformly hyperbolic),
see \cite{DL91} by V.\ Donnay and C.\ Liverani.
So our theorem does not exclude ergodicity for concrete smooth potentials
on~$\bR^d$ and some energies.
But it shows that it would be more difficult to prove ergodicity,
and we would not expect ergodicity for open energy intervals.
For $d\ge 2$ we consider the flow $\Phi\colon\bR\times P\to P$ on phase space
$P:= \bR^d_p\times \bR^d_q$ generated by the natural Hamiltonian function
\[H\in C^2(P,\bR)\qtext,H(p,q)=\eh\norm p^2+V(q)\text,\]
where~$V$ and its first and second
derivatives are assumed to be bounded.
This in particular ensures $\Phi\in C^2(\bR\times P, P)$.
We call the flow $\Phi_E\colon\bR\times \Sigma_E\to\Sigma_E$ restricted
to an energy surface $\Sigma_E:= H^{-1}(E)$ \emph{Anosov} if
there is a $d\Phi_E$--invariant splitting
\nomenclature[Ad]{$d$}{exterior derivative}{}{}%
\[T_x\Sigma_E = {\rm span} \bigl(X_H(x)\bigr) \oplus E^u(x)\oplus E^s(x)\]
\nomenclature[AT]{$T$}{tangential functor}{}{}%
\nomenclature[AXH]{$X_H$}{hamiltonian vectorfield}{}{}%
\nomenclature[AEu]{$E^u$}{strong unstable bundle}{}{}%
\nomenclature[AEs]{$E^s$}{strong stable bundle}{}{}%
into a one-dimensional bundle spanned by the Hamiltonian vector field $X_H$,
and the strong (un)stable bundles $E^{u/s}$, along which $d\Phi_E$
is exponentially contracting in backward resp.\ forward time.
Even if $\Sigma_E$ is not compact, this is unambiguously defined
by existence of $C\ge1$ and $\lambda>0$ with
\begin{equation*}
\norm[\Phi_E(t,x)]{(d\Phi_E^t)_x (v)}
\le C\exp(-\lambda t)\norm[x]v
\qquad\bigl(x\in\Sigma_E,v\in E^s(x),t\in[0,\infty)\bigr)\text,
\end{equation*}
\nomenclature[Glambda]{$\lambda$}{contraction exponent}{}{}%
(and analogously for $E^u$) if we take translation--invariant
norms $\norm[x]{{}\cdot{}}$
on the tangent spaces~$T_xP$ in the bundle $TP$.
\begin{example}\label{exa}
For the potential $V(q)=-\eh\norm q^2$ ($q\in\bR^d$), the flow equals
\begin{equation*}
\V{p(t)\{\vec{q}\,}(t)}
=\V{\idty\cosh(t)&\idty\sinh(t)\\
\idty\sinh(t)&\idty\cosh(t)}\V{p(0)\{\vec{q}\,}(0)}
\qquad(t\in\bR)\text.
\end{equation*}
So we can take $\lambda=C=1$, and
$E^{u/s}(x)=\{\V{a\\\pm a}\mid a\in\bR^d\}$.
We see that on~$P$ there is no \emph{topological}
obstruction against the motion generated by~$H$
to be Anosov (but we remark that here $\dim\bigl(E^{u/s}(x)\bigr)=d$).
\end{example}
The following statement follows from specializing a theorem in \cite{PP94}:
\begin{theorem}[G.\ P.\ and M.\ Paternain]
{\em For no value $E< \sup_q V(q)$ the flow $\Phi_E$ on $\Sigma_E$ is
non--wandering and Anosov.}
\end{theorem}
\begin{proof}
\begin{itemize}[$\bullet$]
\item
To be Anosov, $E$~must be a regular value of~$H$ or -- equivalently -- of~$V$,
since otherwise there are points on $\Sigma_E$,
where the Hamiltonian vector field $X_H$ vanishes.
So one assumes that~$E$ is a regular value, that $\Phi_E$ is
non--wandering and Anosov, and derives a contradiction.
\item
Regularity of~$E$ is one of the assumptions of Theorem~3
in \cite{PP94}.
The condition of existence of a
$\Phi_E$--invariant lagrangian subbundle~$E$ of $T\Sigma_E$ is met,
too, by the centre stable bundle, with
$E(x):= {\rm span} \bigl(X_H(x)\bigr) \oplus E^s(x)\subseteq T_x\Sigma_E$.
\item
As a conclusion Theorem 3 in \cite{PP94} states that $E$ trivially intersects
the vertical bundle ${\rm Vert}$, given for $x\in P\cong T^*\bR^d_q$ by
${\rm Vert}(x)={\rm ker}(d\pi_x)$, with the projection
$\pi\colon T^*\bR^d_q\to\bR^d_q$, $(p,q)\mapsto q$.
This implies that $E\ge \sup_q V(q)$.
For otherwise there is a point $x=(p,q)\in \Sigma_E$ with $V(q)=E$.
By regularity of the value~$E$ then $0\neq X_H(x)\in {\rm Vert}(x)$.
\end{itemize}
\end{proof}
\begin{remark}[The Non-Wandering Condition]
\begin{enumerate}[1.]
\item
If one wants to show ergodicity using the Anosov property,
then the non--wandering condition is somewhat natural:
The Liouville measure $\mu_E$ on $\Sigma_E$
is invariant under the flow $\Phi_E$.
For a regular value~$E$ of~$H$, $\mu_E$ is absolutely continuous
w.r.t.\ the riemannian measure.
The flow is called \emph{ergodic} (in the sense of Aaronson, see \cite{Aar97}) if every
$\Phi_E$--invariant measurable subset~$A$ of $\Sigma_E$ is of
measure zero or the complement of a measure zero set.
Under the assumption of ergodicity, the non--wandering set of
$\Phi_E$ equals~$\Sigma_E$.
For assume that $x\in \Sigma_E$ is wandering.
Then by definition there
is an open neighbourhood $U\subseteq \Sigma_E$%
\nomenclature[AU]{$U$}{neighbourhood of $\Sigma_E$}{}{}
of~$x$ and $T>0$, so that $\Phi_E(t, U) \cap U = \emptyset$ if $\abs t\ge T$.
Since $\dim(\Sigma_E)>1$, there is a neighbourhood $W\subseteq U$%
\nomenclature[AW]{$W$}{neighbourhood of~$U$}{}{}
of~$x$ so that $\mu_E\bigl(\bigcup_{t\in [-T,T]} \Phi_E(t, W)\bigr)
< \mu_E\bigl(U\bigr)$.
So both the $\Phi_E$--invariant set $\bigcup_{t\in \bR} \Phi_E(t,W)$
and its complement have positive measures, contradicting ergodicity.
\item
Besides that, there are many alternatives to the
non--wandering condition in the above theorem.
One choice is to assume that the boundary $V^{-1}(E)\subseteq \bR^d_q$
of Hill's
region contains a compact component which is not diffeomorphic to
$S^{d-1}$, or more than one compact component.
Then there exists a closed ({\em brake}) orbit with positive Maslov class,
contradicting the existence of a section of the lagrangian bundle
(see Theorem 2 of \cite{Kn90} and Section 6 of \cite{KK08}).
Another such alternative is the assumption that there exists an $e<E$
such that for every $r>0$ there exists a ball
$B_r(Q)\subseteq \bR^d_q$ with $V{\upharpoonright}_{B_r(Q)}\le e$. Then the
proof of \cref{theo} below can be adapted.
\end{enumerate}
\end{remark}
For large energies we do not need the non--wandering condition.
\begin{theorem}\label{theo}
For no $E> \sup_q V(q)$ the flow $\Phi_E$ on $\Sigma_E$ is Anosov.
\end{theorem}
\begin{proof}
\begin{itemize}[$\bullet$]
\item
For $E>\sup_q V(q)$ we use riemannian geometry.
The metric~$g$ of a riemannian manifold~$(M^d,g)$
defines a connection and thus a canonical decomposition of~$T(TM)$
into a horizontal and a vertical subspace:
\[T_xTM=T_{x,h}TM\oplus T_{x,v}TM\qquad(x\in TM)\text.\]
Both $T_{x,h}TM$ and $T_{x,v}TM$ are canonically isomorphic to the
$d$-dimensional vector space $T_q M$ (for $x\in T_q M$).
For a lagrangian subspace $\lambda\subseteq T_x TM$ which
is transversal to the vertical subspace, there exists a symmetric operator
\begin{equation}\label{OpS}
S\colon T_{x,h}TM\to T_{x,v}TM
\end{equation}
\nomenclature[AS]{$S$}{symmetric operator $T_{x,h}TM\to T_{x,v}TM$}{}{}%
such that the vertical and horizontal component of any vector
$w=w_h+w_v\in \lambda$ obey the relation (see, \emph{e.g.},
Klingenberg \cite[3.2.16~Proposition]{Kli95})
\begin{equation*}
w_v=Sw_h\text.
\end{equation*}
The covariant derivative $\nabla Y(t)$
\nomenclature[GnablaY]{$\nabla$}{covariant derivative}{}{}%
of a stable Jacobi field~$Y(t)$ along
a geodesic trajectory equals $S(t)Y(t)$.
Hence the operator~$S$ satisfies the Riccati equation
\begin{equation}\label{Ric}
S^2=-\nabla S-R_X\text.
\end{equation}
\nomenclature[ARX]{$R_X$}{curvature operator along~$X$}{}{}%
along the geodesic.
In our case we use on $M:=\bR^d_q$ the Jacobi-Maupertuis metric~$g$,
with $g(q):=\bigl(E-V(q)\bigr)\cdot g_{\textrm{Euclid}}(q)$.
Up to a reparametrisation of time~$t$,
the geodesics in this metric coincide with the projection
of the solutions $t\mapsto \Phi_E(t,x)$
of our Hamiltonian equation to configuration space~$M$.
Since $E-V(q)$ is bounded from below and above by positive constants,
the derivative of time reparametrisation is bounded below and above, too.
We denote the geodesic flow by
\[\Psi:\,\bR\times T_1M\to T_1M\text.\]
\nomenclature[GPsi]{$\Psi$}{geodesic flow}{}{}%
By Theorem~3 of \cite{PP94} we can write the lagrangian subbundle~$E$
as the graph of a symmetric operator valued function of the form~\eqref{OpS}.
We integrate the trace of~\eqref{Ric}
over the unit tangent bundle $T_1B_r$
of the ball $B_r=B_r(0)\subseteq \bR^d$ of radius~$r$.
\item
The integral of the covariant derivative is of order
\begin{equation}
\int_{T_1 B_r} \operatorname{trace}(\nabla S)\, dm\, do = \cO(r^{d-1})\text,
\label{surface}
\end{equation}
where we denote by $dm(q)=\sqrt{\det g(q)}dq_1\wedge\ldots\wedge dq_d$%
\nomenclature[Am]{$m$}{riemannian measure}{}{}
the measure on~$M$ and by $do$%
\nomenclature[Ao]{$o$}{measure on $S^{d-1}$}{}{}
the measure on the unit sphere ($\int_{S^{d-1}}do=\operatorname{vol}(S^{d-1})$).
We show \eqref{surface} by reducing it to a term scaling with
the volume of the boundary $\pa B_r$ of the ball.
To this end we decompose the region $T_1 B_r$ of the energy surface into
\[T_1 B_r\; =\;\cS\;\dot{\cup}\;\cB\;\dot{\cup}\;{\cal T}\text,\]
using the maximal time interval $I(x)$%
\nomenclature[AI]{$I$}{time interval}{}{}
containing~$0$ for which the geodesic flow line through~$x$
stays within $T_1B_r$:
\begin{itemize}[-]
\item The \emph{scattering set}\,\footnote{The names should not be taken too
serious, since, \emph{e.g.}, the intersection of a $\Psi$--orbit with $\cS$
can consist of several components.}~%
$\cS:=\bigl\{x\in T_1 B_r\mid I(x)=[T^-(x),T^+(x)]\;\bigr\}$,
\nomenclature[AS]{$\cS$}{scattering set}{}{}%
\item the \emph{bounded set}~%
$\cB:=\bigl\{x\in T_1B_r\mid I(x)=\bR\bigr\}$%
\nomenclature[AB]{$\cB$}{bounded set}{}{}
and
\item the \emph{trapped set}~%
${\cal T}:=T_1 B_r\setminus(\cB\,\cup\,\cS)$%
\nomenclature[AT]{${\cal T}$}{trapped set}{}{}.
\end{itemize}
All three sets are measurable.
\begin{itemize}[-]
\item
The trapped set ${\cal T}$ consists of wandering points
and thus is of measure zero.
\item
The bounded set $\cB$ is $\Psi$--invariant.
So we can use the relation
\begin{equation}
\int_0^T\operatorname{trace}\bigl(\nabla S\circ\Psi(t,x)\bigr)\,dt =
\operatorname{trace}\bigl(S\circ\Psi(T,x)-S(x)\bigr)
\Leq{Hauptsatz}
to show
\begin{align*}
T\int_{\cB}
\operatorname{trace}(\nabla S)\,dm\,do&
=\int_{\cB}\int_0^T
\operatorname{trace}\bigl(\nabla S\circ\Psi(t,x)\bigr)\,dt\,dm\,do\\&
=\int_{\cB}\operatorname{trace}\bigl(S\circ\Psi(T,x)-S(x)\bigr)\,dm\,do
=0\text.
\end{align*}
\item
So the only contribution to~\eqref{surface}
comes from the scattering set~$\cS$.
Every $x\in\cS$ can be uniquely written as
$x=\Psi(t,y)$ with $t\in[0,T^+(y)]$ and $T^-(y)=0$.
Conversely, for the points in
$\cV:=\{y\in \cS\mid T^-(y)=0\}$ all
$\Psi(t,y)$ with $t\in[0,T^+(y)]$ are in $\cS$.
So we rewrite the integral:
\[\int_\cS\operatorname{trace}(\nabla S)\,dm\,do=\int_\cV\int_0^{T^+(y)}
\hspace*{-4mm}\operatorname{trace}\bigl(\nabla S\circ\Psi(t,y)\bigr) J(y)
\,dt\, dy\text.\]
Since the Jacobian $J\colon\cV\to\bR^+$ is bounded above by~$1$,
and $\operatorname{trace}(S)$ is bounded on~$T_1M$,
we obtain \eqref{surface}, reusing \eqref{Hauptsatz}.
\end{itemize}
\item
For the second term on the right hand side of \eqref{Ric},
\[\int_{T_1 B_r} \operatorname{trace}(R_X)\, dm\, do =
\frac{\operatorname{vol} (S^{d-1})}{d}\int_{B_r} {\cal R}(q)\, dm\text,\]
\nomenclature[AR]{$\mathcal R$}{scalar curvature}{}{}%
where ${\cal R}(q)$ denotes the scalar curvature.
If the motion takes place on a two-dimensional plane $M=\bR^2_q$, then
$\int_{B_r}{\cal R}(q)\,dm=\cO(r)$ as a consequence of the Gauss-Bonnet formula.
For dimension $d\ge3$, that equality is wrong in general.
But in our case the Jacobi metric is conformally flat.
Defining the positive function $u\colon M\to\bR^+$ by
$u(q):=\bigl(E-V(q)\bigr)^{(d-2)/4}$, the measure $dm$ on~$M$
equals
$dm = u^{\frac{2d}{d-2}}dq_1\wedge\ldots\wedge dq_d$.
The scalar curvature equals
\[{\cal R}=\frac{1-d}{(E-V)^3} \l[ (V-E)\Delta V + \frac{d-6}{4}
(\nabla V)^2 \ri] = 4\frac{1-d}{d-2}u^{-\frac{d+2}{d-2}}\Delta u\]
(with the euclidean Laplacian $\Delta=\sum_{k=1}^d \frac{\pa^2}{\pa q_k^2}$).
\nomenclature[GDelta]{$\Delta$}{euclidian Laplacian}{}{}%
Therefore
\begin{align}
\lefteqn{\int_{B_r}{\cal R}\,dm
=-4\frac{d-1}{d-2}\int_{B_r}u^{-\frac{d+2}{d-2}}
(\Delta u) u^{\frac{2d}{d-2}}dq_1\wedge\ldots\wedge dq_d}\nonumber\\&
=-4\frac{d-1}{d-2}\int_{B_r} u(\Delta u)
dq_1\wedge\ldots \wedge dq_d\nonumber\\&
=+4\frac{d-1}{d-2}\int_{B_r} (\nabla u)(\nabla u)
dq_1\wedge\ldots \wedge dq_d + f(r)
\ge f(r)\text.\label{larger}
\end{align}
The surface integral is of order $f(r)=\cO(r^{d-1})$.
Concerning the left hand side of \eqref{Ric},
the integral of $\operatorname{trace}(S^2)$ is positive.
For uniform hyperbolicity, one would need
$\limsup_{r\to\infty}r^{-d}\int_{B_r}\operatorname{trace}(S^2)(q)\,dm>0$,
which is impossible, since the corresponding $\limsup$ of the right hand side is
nonpositive.
So \eqref{larger} is compatible with \eqref{Ric} only if
the flow is not Anosov.
\end{itemize}
\end{proof}
\section{Random Coulombic Potentials}\label{sec9}
As it does not seem to be so simple to find smooth random potentials that
lead to ergodic motion, we now study the example of coulombic potentials,
see \cref{fig:random} for a numerical realization
\begin{figure}
\begin{center}
\ifpdf
\includegraphics[width=5cm,clip]{simulation3.pdf}
\else
\includegraphics[width=5cm,clip]{simulation3.ps}
\fi
\end{center}
\caption{Motion in the configuration space of a random coulombic potential}
\label{fig:random}
\end{figure}
We restrict ourselves to dimension $d=2$.
Here for $j\in J$ the single site potentials
\[\tilde W_j\in C^\eta(\bR^2\setminus\{s_j\},\bR)\]
\nomenclature[AWtildej]{$\tilde W_j$}{single site potential with singularity}{}{}%
(with $\eta\in\bN\cup\{\infty\},\eta\ge2$)
diverge at the position $s_j\in \cD$ in the fundamental domain
$\cD:=\{x_1\ell_1+x_2\ell_2\mid 0\leq x_i<1\}$
of the lattice ${\mathcal L}=\Span_\bZ(\ell_1,\ell_2)$.
Our assumptions are:
\begin{compactenum}
\item\label{shortrange}
\emph{The decay at infinity} is short range,
that is for $\alpha\in\bN_0^2,\ \abs\alpha\leq\eta$
\[\pa^\alpha\tilde W_j(q)=\cO(\norm q^{-2-\vep})\qquad(\norm q\rightarrow\infty)\text.\]
\item\label{coulomb}
The \emph{local singularity} at $s_j\in\bR^2\cong\bC$ is controlled by
\[f_j\colon\bC^*:=\bC\setminus\{0\}\to\bR \qtextq{,}
f_j(z):=\abs z^2\tilde W_j(z^2+s_j)\]
and we assume that for all multi--indices $\alpha\in\bN_0^2$,
$\abs\alpha\leq\eta$, $\pa^\alpha f_j$
can be continuously extended to zero, with $f_j(0)<0$.
We allow for an additional single site potential $\tilde W_0=0$.
\end{compactenum}
\begin{example
\label{ex:Yukawa}
The \emph{Yukawa Potential} with parameters $c_j,\mu_j>0$
is defined via
\[\tilde W_j(q)=-c_j\frac{\exp(-\mu_j\norm{q-s_j})}{\norm{q-s_j}}\text.\]
\end{example}
\begin{example
\label{ex:finiteRange}
\emph{Finite range potentials}
are given by
\[\tilde W_j(q):=-\frac{g_j(\norm{q-s_j})}{\norm{q-s_j}}\text,\]
with $g_j\in C_c^\eta(\bR,\bR),\ g_j(0)>0$.
\nomenclature[Agj]{$g_j$}{radial shape of finite range potential}{}{}%
\end{example}
The random potential is determined by the probability space
$(\Omega,\cB(\Omega),\beta)$ with $\Omega={\mathcal L}^J$.
The probability measure~$\beta$ is assumed to be $\vartheta$--invariant,
see equation~\eqref{eq:vartheta},
and to give probability $\beta\{\omega_0\}=0$
to the configuration $\omega_0\in\Omega$ with $\omega_0(\ell)=0$, $\ell\in{\mathcal L}$.
No ${\mathcal L}$--ergodicity of $\beta$ is assumed here.
For what follows we fix $\omega\in\Omega$.
The \emph{punctured configuration space}
$\tilde M_\omega:=\bC\setminus\cS_\omega$
\nomenclature[AMtildeomega]{$\tilde M_\omega$}{punctured configuration space}{}{}%
now depends on the \emph{singularity set}
$\cS_\omega:=\{s_{\omega(\ell)}+\ell\mid\ell\in{\mathcal L},\omega(\ell)\ne0\}$.
\nomenclature[ASomega]{$\cS_\omega$}{singularity set}{}{}%
The former supports the random potential
\begin{equation*}
\tilde V_\omega\colon\tilde M_\omega\to\bR\qtextq,
\tilde V_\omega(q):=\sum_{\ell\in{\mathcal L}}\tilde W_{\omega(\ell)}(q-\ell)\text.
\end{equation*}
\nomenclature[AVtildeomega]{$\tilde V_\omega$}{coulombic potential}{}{}%
Assumption~\ref{shortrange}.\ guarantees the convergence of $\tilde V_\omega$
and its derivatives,
and Assumption~\ref{coulomb}.\ implies that the Coulombic singularities
are attractive and
\begin{equation*}
\tilde V_{\omega,\max}:=\sup\tilde V_\omega(\tilde M_\omega)<\infty\text.
\end{equation*}
\nomenclature[AVtildeomegamax]{$\tilde V_{\omega,\max}$}{supremum of $\tV_\omega$}{}{}%
The Hamiltonian flow $\tilde\Phi_\omega\colon\tilde U_\omega\to\tilde P_\omega$
\nomenclature[GPhitildeomega]{$\tilde Phi_\omega$}{incomplete hamiltonian flow}{}{}%
on $\tilde P_\omega:=T^*\tilde M_\omega=\bR^2\times\tilde M_\omega$,
generated by the Hamiltonian function
$\tilde H_\omega\colon\tilde P\to\bR$, $(p,q)=\frac12\norm p^2+\tilde V_\omega(q)$,
\nomenclature[GHtildeomega]{$\tilde H_\omega$}{coulombic hamiltonian function}{}{}%
is now, due to the Coulombic singularities, incomplete and only defined
on a maximal open subset $\tilde U_\omega\subseteq\bR\times\tilde P_\omega$.
\nomenclature[AUtildeomega]{$\tilde U_\omega$}{maximal domain of $\tilde Phi$}{}{}%
It is known that the flow can be continuously
regularised by reflecting collision orbits at their singularity.
This is possible, see Prop.\ 2.3 of \cite{KK92} or
Thm.~11.23 of \cite{Kn11},
by smoothly extending the incomplete Hamiltonian system
\[\big(\tilde P_\omega,dq\wedge dp,\tilde H_\omega\big)
\quad\text{ to a Hamiltonian system }\quad
\big(P_\omega, \sigma_\omega, H_\omega\big)\text,\]
\nomenclature[APomegasigmaomega]{$(P_\omega,\sigma_\omega)$}{symplectic manifold}{}{}%
\nomenclature[Gsigmaomega]{$\sigma_\omega$}{symplectic form}{}{}%
with $H_\omega\colon P_\omega\to \bR$
generating a complete smooth Hamiltonian flow
on the symplectic manifold $(P_\omega, \sigma_\omega)$.
In fact, similarly to the construction in \cref{sec2},
these data extend continuously to a triple $(P,\sigma,H)$, with
extended phase space
$P:= \bigcup_{\omega\in\Omega}\ \{\omega\}\times P_\omega$.
For $d=2$ and energies $E>V_{\max}$ regularization can also be performed,
as in \cite{Kna87,KK92}, with the help of the twofold covering
\begin{equation}\label{eq:Momega}
\pi_\omega\colon\Mo:= \big\{(q,Q)\in\bC^2\mid f_\omega(q)=Q^2\big\}
\; \longrightarrow \;\bC
\end{equation}
\nomenclature[Gpiomega]{$\pi_\omega$}{twofold covering $\Mo\to\bC$}{}{}%
\nomenclature[AMomega]{$\Mo$}{branched covering surface}{}{}%
with branch points in the singularity set
$\cS_\omega$, where $f_\omega\colon\bC\to\bC$
\nomenclature[AMomega]{$f_\omega$}{entire function}{}{}%
is an entire function with simple zeroes in and only in~$\cS_\omega$:
$f_\omega^{-1}\{0\}=\cS_\omega$.
By a Weierstrass product construction
we can choose $(\omega,q)\mapsto f_\omega(q)$
as a continuous function.
The lift $\mathbf{\tilde g}_{\omega,E}:=(\pi_\omega)^*\tilde g_{\omega,E}$
\nomenclature[AgtildeomegaE]{$\mathbf{\hat g}_{\omega,E}$}{Jacobi metric on $\Mo$ without branch points}{}{}%
of the Jacobi-Maupertius metric
\begin{equation*}
\tilde g_{\omega,E}(q):=\bigl(1-E^{-1}\tilde V_\omega(q)\bigr)g_{\Euclid}(q)
\qquad(q\in\tM_\omega)
\end{equation*}
can be continued to all of $\Mo$ by taking limits:
\begin{equation*}
\mathbf g_{\omega,E}(q,0):=
\lim\limits_{\Mo\ni(q',Q)\to(q,0)}\mathbf{\hat g}_{\omega,E}(q',Q)\qquad
(\pi_\omega(q)\in\cS_\omega)\text,
\end{equation*}
\nomenclature[AgomegaE]{$\mathbf g_{\omega,E}$}{Jacobi metric on $\Mo$}{}{}%
see \cite{Kna87,KK92}.
This gives a smooth and complete riemannian metric on~$\Mo$,
and its geodesics are, up to a reparametrisation of time
and modulo~$\pi_\omega$, trajectories of the Hamiltonian flow.
This geometric regularisation allows to take full advantage of riemannian geometry.
For two-dimensional compact riemannian manifolds of negative curvature
ergodicity of the geodesic flow was established by \cite{Hop41}.
The gaussian curvature of $(\tM_\omega,\hat g_{\omega,E})$
at $q\in\tM_\omega$ is given by
\begin{equation}\label{eq:curvature}
\begin{split}
\tilde K_{\omega,E}(q)&
=\frac{\bigl(E-\tilde V_\omega(q)\bigr)\Delta\tilde V_\omega(q)
+\bigl(\nabla\tilde V_\omega(q)\bigr)^2}
{2\bigl(E-\tilde V_\omega(q)\bigr)^3}
\\&
=\frac{\bigl(1-E^{-1}\tilde V_\omega(q)\bigr)E^{-1}
\Delta\tilde V_\omega(q)
+\bigl(E^{-1}\nabla\tilde V_\omega(q)\bigr)^2}
{2E\cdot\bigl(1-E^{-1}\tilde V_\omega(q)\bigr)^3}
\text.
\end{split}
\end{equation}
\nomenclature[AKtildeomegaE]{$\tilde K_{\omega,E}$}{gaussian curvature}{}{}%
Since $\pi_\omega^{-1}(\tilde M_\omega)$ is a local isometry,
the curvature of all non--branch points of~$\Mo$
is determined by \cref{eq:curvature}.
In \cref{ex:Yukawa}
it turns out that for high enough energy
$E>E_{\text{th}}>\tilde V_{\omega,\max}$ the curvature is non--positive.
This motivates the following definition.
\begin{defi}
The pair $(\tilde V_\omega,E)$ is of
\emph{non--positive (strictly negative) curvature},
if $(\Mo,\mathbf g_{\omega,E})$
has non--positive (strictly negative) curvature.
\end{defi}
A direct consequence of equation~\eqref{eq:curvature} is, that,
given a pair $(\tilde V_\omega,E)$ of non--positive curvature,
all $(\tilde V_\omega,E')$ with $E'\ge E$
are of non--positive curvature, too.
As $E\nearrow\infty$ the curvature concentrates in the singularities.
In \cref{ex:finiteRange}
non--positive curvature can be achieved
for all energies above a threshold with suitable choices for $g_j$, \emph{e.g.}
\begin{equation*
g_j(r):=
\begin{cases}
-c_j\cos^{\eta+1}(\lambda_jr)
&r<\frac\pi{2\lambda}\\
0 &r\ge\frac\pi{2\lambda}
\end{cases}
\end{equation*}
with $c_j,\lambda_j>0$, see \cite{Kna87}. From now on we will assume non--positive curvature.
This assumption and the following lemma explain why we allow only $W_0=0$
as smooth single site potential.
\begin{lemma}
If a single site potential $\tilde W\colon\bC\to\bR$
without singularity and an energy~$E$
are of nonpositive curvature, then $\tilde W=\tilde W_0\equiv0$.
\end{lemma}
\begin{proof}
From \eqref{as:sisi} it follows that the curvature of one single site potential is integrable.
By the theorem of Gauss--Bonnet
the integral of the curvature over~$\bC$ vanishes.
Since the curvature is nonpositive, it has to vanish, too,
and so does the single site potential.
\end{proof}
Our goal is
\begin{theorem}\label{thm:toptrans
If $\,(\tilde V_\omega,E)$ is of non--positive curvature, then
the geodesic flow
$\Phi_\omega\colon\bR\times T_1\Mo\to T_1\Mo$
\nomenclature[AT1]{$T_1$}{unit tangent bundle functor}{}{}%
is \emph{topologically transitive} $\beta$--almost surely.
\end{theorem}
The strategy will be as follows.
In \cref{prop:periodicdense} we show that
the periodic orbits of~$\Phi_\omega$ are dense in~$T_1\Mo$.
Note, that the same is true for Anosov diffeomorphisms on compact manifolds,
cf.~\cite[3.8]{Bow75}.
Then, in the proof of \cref{thm:toptrans} from page~\pageref{proof-toptrans} on,
in order to connect two open subsets of $T_1\Mo$,
we will connect two periodic orbits with an intertwining orbit.
\newcommand{\vphi}{\varphi}%
\newcommand{\cone}[3]{\vartriangle_{#1}^{(#2)}\!(#3)}%
\newcommand{\sector}[3]{\blacktriangle_{#1}^{(#2)}(#3)}%
We start with a useful lemma, which singles out the set of full measure
on which we establish topological transitivity.
We denote the Euclidean cone intersected with the lattice~${\mathcal L}$
\begin{equation*}
\cone q\vphi x:=\{\ell\in{\mathcal L}\mid\sphericalangle(x,\ell-q)<\vphi\}\text,
\end{equation*}
\nomenclature[Gdelta]{$\cone q\vphi x$}{euclidian cone}{}{}%
$q\in\bC$, $\vphi\in[0,\pi]$, $x\in T_{1,q}\bC=S^1$.
\begin{lemma}\label{lemma:nofreecones}
The $\vartheta$--invariance of $\beta$ and $\beta\{\omega_0\}=0$
imply that there are $\beta$-almost surely no Euclidean cones
without nonvanishing single site potential.
More precisely:
\begin{equation*}
\beta\bigl\{\omega\in\Omega\bigm|
\exists(q,\vphi,x)\in\bR^2\times(0,\pi]\times
S^{1}\colon
\omega(\cone q\vphi x)=\{0\}\bigr\}=0\text
\end{equation*}
\end{lemma}
\begin{proof}
We study the sets
\begin{equation*}
A_{q,\vphi,x}:=\bigl\{\omega\in\Omega\bigm|
\omega\bigl(\cone q\vphi x\bigr) = \{0\} \bigr\}\text,
\end{equation*}
$(q,\vphi,x)\in\bR^{2}\times(0,\pi]\times
S^{1}$, first.
For $(q,\vphi,\ell)\in\bR^2\times(0,\pi]\times{\mathcal L}$ we have
\begin{equation*}
\beta\bigl(A_{q,\vphi,\frac\ell{\norm\ell}}\bigr)
=\lim_{n\to\infty}\beta\bigl(A_{q-n\ell,\vphi,\frac\ell{\norm\ell}}\bigr)
=\beta\Bigl(\Isect\nolimits_{n\in\bN}A_{q-n\ell,\vphi,\frac\ell{\norm\ell}}\Bigr)
=\beta\{\omega_0\}
=0\text.
\end{equation*}
Note now that the set in question can be written as
the denumerable union
\begin{equation*}
\Union\nolimits_{(q,\vphi,\ell)\in\bQ^2\times\bQ_+\times{\mathcal L}}
A_{q,\vphi,\frac\ell{\norm\ell}}\text.
\end{equation*}
\end{proof}
We denote the tangent bundle of $\Mo$ with $\tau_{\Mo}\colon T\Mo\to\Mo$
\nomenclature[GtauMomega]{$\tau_{\Mo}$}{tangential projection}{}{}%
and the natural length metric on $(\Mo,\mathbf g_{\omega,E})$ with $d_{\Mo}$.
\nomenclature[AdMomega]{$d_{\Mo}$}{natural length metric on $\Mo$}{}{}%
\begin{lemma}\label{lemma:singsappear}
For $\beta$--almost all $\omega\in\Omega$ the following holds.
Given a point $q\in\Mo$ and a nonempty open subset
$U\subseteq T_{1,q}\Mo:=T_1\Mo\isect T_q\Mo$ of the sphere bundle,
there exists an initial direction $v\in U$
such that the corresponding geodesic hits a branch point at $t>0$, i.e.
\begin{equation*}
\exp_{\Mo}(tv)\in\pi_\omega^{-1}(\cS_\omega)\text.
\end{equation*}
\end{lemma}
\begin{proof}
W.l.o.g.\ we assume $U\ne T_{1,q}\Mo$.
For every $\omega\in\Omega$ such that there is an open subset
$U\subseteq T_{1,q}\Mo$ with no branch points in
\begin{equation*}
\blacktriangle:=\exp_{\Mo}(\bR_+U)\;\subseteq\;\Mo
\end{equation*}
\nomenclature[Gdelta]{$\blacktriangle$}{Jacobian cone}{}{}%
we will establish the existence of a Euclidean cone
$\vartriangle\subseteq{\mathcal L}$ with $\omega(\vartriangle)=\{0\}$.
Then \cref{lemma:nofreecones} applies and gives the desired result.
\par
Fix $v\in U$.
Since the curvature of $(\Mo,\mathbf g_{\omega,E})$ is nonpositive,
we find a constant $\rho_0>0$ such that
\begin{equation}\label{eq:bigdistance}
d_{\Mo}\bigl(\exp_{\Mo}(tv),\,\Mo\setminus\blacktriangle\bigr)
\;\ge\;2\rho_0t
\end{equation}
for all $t\ge0$, see \cite{BBI01}.
Note that, due to the absence of branch points in~$\blacktriangle$,
$\pi_\omega|_{\blacktriangle}$ is a homeomorphism onto its image,
and since the conformal factor is bounded by
$h:=1-E^{-1}\tilde V_{\omega,\max}$, we know for all $q,q'\in\Mo$
\begin{equation*}
d_{\Mo}(q,q')
\;\le\; h \cdot\abs{\pi_\omega(q)-\pi_\omega(q')}\text.
\end{equation*}
This implies
\begin{equation*}
\blacktriangle'
\;:=\;\Union_{t\ge0}B_{h^{-1}\rho_0t}
\bigl(\pi_\omega\circ\exp_{\Mo}(tv)\bigr)
\;\subseteq\;\pi_\omega(\blacktriangle)
\;\subseteq\;\bC\text,
\end{equation*}
and in $\blacktriangle'$ we search our Euclidean cone~$\vartriangle$.
All we have to show to this end is that the ``axis'' of $\blacktriangle'$
\begin{equation*}
\bR_+\ni t\;\longmapsto\;\bigl(q_\omega(t),p_\omega(t)\bigr)
:=T\pi_\omega\circ\Phi_\omega(t,v)
\end{equation*}
converges to a definite direction:
$ \lim_{t_0\to\infty}
\sup_{t\ge t_0} \abs{p_\omega(t+t_0,v)-p_\omega(t_0,v)}=0$.
But this is clear from Assumption~1,
\emph{i.e.}\ that all single site potentials are short range,
and equation~\eqref{eq:bigdistance}, which together imply
that the curvature along the axis vanishes.
\end{proof}
With this tools at hand it is easy to prove
\begin{proposition}\label{prop:periodicdense}
The set of periodic orbits of $\Phi_\omega\colon\bR\times T_1\Mo\to T_1\Mo$
is dense in $T_1\Mo$ for $\beta$--almost every $\omega\in\Omega$.
\end{proposition}
\begin{proof}
To any nonempty open set $U\subseteq T_1\Mo$
\nomenclature[AU]{$U$}{open subset of $T_1\Mo$}{}{}%
we construct a periodic orbit whose trajectory intersects~$U$.
\Cref{lemma:singsappear} guarantees that we hit
a branch point at some time~$t>0$.
We choose a branch point
\begin{equation*}
x\;\in\;\exp_{\Mo}(tU)
\isect\pi_\omega^{-1}(\cS_\omega)\;\subseteq\;\Mo\text.
\end{equation*}
Now $\Phi_\omega$ is continuous, which shows that
\begin{equation*}
U':=-\Phi_\omega(t_1,U)\isect T_{1,x}\Mo
\end{equation*}
is open in $T_{1,x}\Mo$.
Again by \cref{lemma:singsappear} we find $v\in U'$ and $t'>0$
such that $\Phi_\omega(t',v)$ hits again a branch point.
By construction the trajectory of~$v$ intersects~$U$ and is periodic.
\end{proof}
To gain more overview we introduce the universal covering
$\mathbf\pi_\omega^*\colon\Mo^*\to\Mo$ of $\Mo$
\nomenclature[Gpistaromega]{$\mathbf\pi_\omega^*$}{universal covering of~$\Mo$}{}{}%
\nomenclature[AMstaromega]{$\Mo^*$}{universal covering surface of~$\Mo$}{}{}%
and equipp it with the riemannian metric
$\mathbf g_{\omega,E}^*:=(\mathbf\pi_\omega^*)^*\mathbf g_{\omega,E}$.
\nomenclature[AgstaromegaE]{$\mathbf g_{\omega,E}^*$}{Jacobi metric on~$\Mo^*$}{}{}%
This makes $\Mo^*$~a Hadamard manifold.
We denote the natural length metric with $d_{\Mo^*}$.
\nomenclature[Adstaromega]{$d_{\Mo^*}$}{natural length on~$\Mo^*$}{}{}%
The fact that there is a certain amount of negative curvature in~$\Mo^*$
is expressed in the following.
\begin{proposition}\label{prop:visibility}
For $\beta$--almost all $\omega\in\Omega$
the riemannian surface $(\Mo^*,\mathbf g_\omega^*)$
is a \emph{visibility manifold}, i.e.\ for all $p\in\Mo^*$
and $\ve>0$ there exists $r=r(p,\ve)>0$ such that all geodesic segments
\nomenclature[Gsigma]{$\sigma$}{geodesic segment}{}{}%
$\sigma\colon[a,b]\to\Mo^*$ with distance at least~$r$ from~$p$
are seen from~$p$ under an angle less then~$\ve$ (cf.\ \cite{EO73}):
\begin{equation*}
\forall p\in\Mo^*,\ve>0\exists r>0\colon\quad
d_{\Mo^*}(p,\sigma[a,b])\ge r
\quad\implies\quad
\sphericalangle_p\bigl(\sigma(a),\sigma(b)\bigr)<\ve\text.
\end{equation*}
\nomenclature[Gangelp]{$\sphericalangle_p$}{angel in $p$}{}{}%
\end{proposition}
\begin{proof}
We need to show that the set of all $\omega\in\Omega$
which allow a positive~$\ve$ such that we find for all $r>0$
a geodesic segment $\sigma_r\colon[a_r,b_r]\to\Mo^*$ with
\begin{inparaenum}
\item $d_{\Mo^*}(p,\sigma_r)>r$ and
\item $\sphericalangle_p\bigl(\sigma_r(a_r),\sigma_r(b_r)\bigr)\ge\ve$
\end{inparaenum}
is of $\beta$--measure~$0$.
To do this, we find a cone without singularities
and invoke \cref{lemma:singsappear}.
\par
By compactness of $T_{1,p}\Mo^*$
we get $x,y\in T_{1,p}\Mo^*$
with $\sphericalangle_p(x,y)>\ve$ and
\begin{equation*}
d_{\Mo^*}(p,\gamma_t)\xto{t\to\infty}\infty\text,
\end{equation*}
where $\gamma_t\colon[a_t,b_t]\to\Mo^*$
is the unique geodesic segment connecting
$\gamma_t(a_t)=\exp_{\Mo^*}(tx)$ and
$\gamma_t(b_t)=\exp_{\Mo^*}(ty)$.\\
%
The theorem of Gauss--Bonnet implies that the curvature
integrated over the cone
\begin{equation*}
\blacktriangle:=
\Union_{t>0}\gamma_t[a_t,b_t]
\end{equation*}
between~$x$ and~$y$ is bounded by~$-\pi$ from below.
This means that $\blacktriangle$ cannot cover two singularities,
since then $\blacktriangle$, beeing geodesically convex,
would cover the periodic orbit connecting these two,
and this would contradict Gauss--Bonnet.
Thereby $\blacktriangle$ contains a cone without singularity
and \cref{lemma:singsappear} applies.
\end{proof}
\begin{remark}\label{rem:visibility}
Two geodesic rays $\gamma,\gamma'\colon\bR_+\to\Mo^*$
are called \emph{asymptotic}, if their distance is bounded.
Being asymptotic is an equivalence relation,
and the set of equivalence classes
(equipped with a suitable topology, see \cite{EO73})
is the \emph{ideal boundary} $\partial\Mo^*\equiv\Mo^*(\infty)$
\nomenclature[AMhatomegastar]{$\partial\Mo^*\equiv\Mo^*(\infty)$}{ideal boundary of $\Mo^*$}{}{}%
of $\Mo^*$.
A point $\gamma(\infty):=[\gamma]\in\Mo^*(\infty)$
\nomenclature[Ggammainfty]{$\gamma(\infty)$}{end point of~$\gamma$}{}{}%
is a \emph{zero point}, if for all $\gamma'\in\gamma(\infty)$
the distance between $\gamma$ and $\gamma'$ vanishes.
\cite{EO73} explains that a visibility manifold
satisfies \emph{Axiom~1}, i.e.\ any two distinct boundary points
can be connected with a (not necessarily unique) geodesic.
\end{remark}
\begin{lemma}\label{lemma:zeropoint}
For all $\omega\in\Omega$ holds the following.
Every lift of a periodic geodesic in $\Mo$ to $\Mo^*$
connects two zero points of $\Mo^*(\infty)$.
\end{lemma}
\begin{proof}
Given a periodic geodesic $\gamma\colon S^1\to\Mo$,
its lift $\gamma^*\colon\bR\to\Mo^*$, and another geodesic
$\gamma'\colon\bR\to\Mo^*$ with $\gamma^*(\infty)=\gamma'(\infty)$,
we integrate curvature over the area bounded by $\gamma^*(\bR_+)$,
$\gamma'(\bR_+)$ and the geodesic segment connecting $\gamma^*(0)$
and $\gamma'(0)$.
Gauss--Bonnet assures us, that this quantity is bounded,
and this is only possible if $\gamma^*$ and $\gamma'$ have distance~$0$.
Otherwise, since $\gamma^*$ covers a periodic orbit, the curvature
integral is unbounded.
\end{proof}
\begin{proof}[of \cref{thm:toptrans}]\label{proof-toptrans}
\Cref{prop:periodicdense} tells us that
for almost all $\omega\in\Omega$ periodic orbits of $\Phi_\omega$
are dense in $T_1\Mo$.
Therefore for any two open and nonempty sets
$U,V\subseteq T_1\Mo$ we find
two periodic geodesics $\gamma_U,\gamma_V\colon\bR\to\Mo$
which intersect~$U$ and~$V$, respectively,
i.e.\ $\dot\gamma_U(\bR)\isect U\ne\emptyset$ and $\dot\gamma_U(\bR)\isect V\ne\emptyset$.
\par
There are lifts $\gamma_U^*,\gamma_V^*\colon\bR\to\Mo^*$
of $\gamma_U$ and $\gamma_V$ to the universal covering
$\mathbf\pi_\omega^*\colon\Mo^*\to M_\omega$ of $\Mo$.
By \cref{prop:visibility,rem:visibility}
the endpoints $\gamma_U^*(-\infty)$ and $\gamma_V^*(\infty)$
can be joined with a geodesic $\gamma^*\colon\bR\to\Mo^*$,
i.e.\ $\gamma^*(-\infty)=\gamma_U^*(-\infty)$ and
$\gamma^*(\infty)=\gamma_V^*(\infty)$.
\Cref{lemma:zeropoint} makes sure that the distance between
$\gamma^*(t)$ and $\gamma_U^*\bigl((0,\infty)\bigr)$
respectively $\gamma_V^*\bigl((-\infty,0)\bigr)$ vanishes,
as $t\to\pm\infty$, respectively:
\begin{equation*}
\lim_{t\to\infty}d_{\Mo^*}
\bigl(\gamma^*(t),\gamma_U^*\bigl((0,\infty)\bigr)\bigr)
=\lim_{t\to-\infty}d_{\Mo^*}
\bigl(\gamma^*(t),\gamma_V^*\bigl((-\infty,0)\bigr)\bigr)
=0\text.
\end{equation*}
This implies that $T\mathbf\pi_\omega^*\circ\dot\gamma^*(\bR)$
intersects~$U$ and~$V$.
\end{proof}
\section{Ergodicity of the finite factor}\label{sect10}
Similar to the lattice and the poissonian case,
the Coulombic system does have a finite factor.
Assuming $(\vartheta,\beta)$ to be ergodic again
and the single site potentials to be $\eta\ge3$
times continuously differentiable, we
show ergodicity of this factor.
This is analogous to a result in \cite{Len03} in the setting of billiards.
Thanks to the smoothness of our system,
our proof can rely directly on \cite{Hop41}
without the need for technical generalisations
to cope with singularities like \emph{e.g.} in \cite{LW95}.
The following construction of the finite factor
is carried out in detail in \cite{Sch04,Sch10}.
As above the geodesic motion on the energy surface
is regularised with the twofold covering~\eqref{eq:Momega}.
There is one nontrivial deck transformation
\begin{equation*}
\mathbf G_\omega\colon\mathbf M_\omega\to M_\omega\textq,
G(q,Q):=(q,-Q)\text.
\end{equation*}
\nomenclature[AGomega]{$\mathbf G_\omega$}{deck transformation}{}{}%
For each $\omega\in\Omega$ and every $\ell\in{\mathcal L}$
there exists an isometry
\begin{equation*}
\mathbf\phi_{\omega,\ell}\colon
\mathbf M_\omega\to\mathbf M_{\vartheta_\ell\omega}
\end{equation*}
\nomenclature[Gphiomegal]{$\mathbf\phi_{\omega,\ell}$}{isometry $\mathbf M_\omega\to\mathbf M_{\vartheta_\ell\omega}$}{}{}%
that shifts $q$ to $q-\ell$:
\begin{equation*}
\pi_{\vartheta_\ell\omega}\circ\mathbf\phi_{\omega,\ell}(q,Q)
\;=\;q-\ell\qquad\bigl((q,Q)\in\mathbf M_\omega\bigr)\text.
\end{equation*}
The deck transformation gives another such map
$\mathbf\phi_{\omega,\ell}\circ\mathbf G_\omega
=\mathbf G_{\vartheta_\ell\omega}\circ\mathbf\phi_{\omega,\ell}$.
Before we can bundle the maps
$\bigl((\mathbf\phi_{\omega,\ell})_{\omega\in\Omega}\bigr)_{\ell\in{\mathcal L}}$
into a group action of~${\mathcal L}$
on~$\mathbf M:=\Union_{\omega\in\Omega}\{\omega\}\times\mathbf M_\omega$,
\nomenclature[AM]{$\mathbf M$}{twofold branched covering of extended phase space}{}{}%
we have to divide through the group of deck transformations.
But $\pi_\omega$ is a branched covering,
so to keep the differentiable structure we first follow \cite{Kna87,KK92}
and restrict the geodesic flow to the energy surface
\begin{equation*}
\mathbf\Sigma_{\omega,E}
:=\big\{(x,v)\in T\bC^2\mid\mathbf g_{\omega,E}(v,v)=1\big\}\text.
\end{equation*}
\nomenclature[GSigmaomegaE]{$\mathbf\Sigma_{\omega,E}$}{unit sphere bundle of branched covering of $\mathbf M_\omega$}{}{}%
We divide $\mathbf\Sigma_{\omega,E}$ by the group~%
$\bZ_2\cong\big\{\Id_{\mathbf\Sigma_{\omega,E}},
D\mathbf G_\omega{\upharpoonright}_{\mathbf\Sigma_{\omega,E}}\big\}$
and get the smooth manifold
\begin{equation*}
\Sigma_{\omega,E}:=\mathbf\Sigma_{\omega,E}/\bZ_2\text.
\end{equation*}
\nomenclature[GSigmaomegaE]{$\Sigma_{\omega,E}$}{$\bZ_2$--quotient of $\mathbf\Sigma_{\omega,E}$}{}{}%
The disjoint union
\begin{equation*}
\mathbf\Sigma_E
:=\Union_{\omega\in\Omega}\{\omega\}\times\mathbf\Sigma_{\omega,E}
\end{equation*}
\nomenclature[GSigmaE]{$\mathbf\Sigma_E$}{$\bZ_2$--quotient of $\mathbf\Sigma_E$}{}{}%
inherits its topology from the embedding into $\Omega\times T\bC^2$.
The quotient
\begin{equation*}
\Sigma_E:=\mathbf\Sigma_E/\bZ_2
=\Union_{\omega\in\Omega}\{\omega\}\times\Sigma_{\omega,E}
\end{equation*}
is the phase space for which we can construct the desired group action
\begin{equation*}
\Theta^E\colon{\mathcal L}\times\Sigma_E\to\Sigma_E\textq,
\Theta^E_\ell\bigl(\omega,[x]_{\bZ_2}\bigr)
:=\bigl(\vartheta_\ell\omega,[D\mathbf\phi_{\omega,\ell}(x)]_{\bZ_2}\bigr)\text.
\end{equation*}
\nomenclature[GThetaE]{$\Theta^E$}{${\mathcal L}$--action on $\Sigma_E$}{}{}%
This group action is continuous.
Analogous to \cref{sec2} the quotient
\begin{equation*}
\pi_{\mathcal L}\colon\Sigma_E\to\hat\Sigma_E:=\Sigma_E/{\mathcal L}
\end{equation*}
\nomenclature[GpiL]{$\pi_{\mathcal L}$}{quotient $\Sigma_E\to\widetilde\Sigma_E$}{}{}%
\nomenclature[GSigmahatE]{$\hat\Sigma_E$}{compact coulomb phase space}{}{}%
is compact and metrizable, inherits the geodesic flow
\begin{equation*}
\hat\Phi_E\colon\bR\times\hat\Sigma_E\to\hat\Sigma_E
\end{equation*}
\nomenclature[GPhihatE]{$\hat\Phi_E$}{geodesic flow on $\hat\Sigma_E$}{}{}%
and carries the finite and $\hat\Phi_E$--invariant
Liouville measure~$\hat\mu_E$, see \cite{Sch10}.
\nomenclature[GMstar]{$\hat\mu_E$}{Liouville measure on $\hat\Sigma_E$}{}{}%
An overview over the different phase spaces including
the embedding into $\Omega\times T\bC^2$
and the unit tangent bundle~$\mathbf\Sigma_{\omega,E}^*$
\nomenclature[GSigmastaromegaE]{$\mathbf\Sigma_{\omega,E}^*$}{unit tangent bundle of $\Mo^*$}{}{}%
of the universal cover~$\mathbf\pi_\omega^*\colon\Mo^*\to\Mo$
is given in \cref{abb:orientierung}.
\nomenclature[GPistaromega]{$\mathbf\Pi_\omega^*$}{covering $\mathbf\Sigma_{\omega,E}^*\to\mathbf\Sigma_{\omega,E}$}{}{}%
\begin{figure}[h
\begin{equation*}
\begin{CD}
\mathbf\Sigma_{\omega,E}^*@>\mathbf\Pi_\omega^*>>
\mathbf\Sigma_{\omega,E}@>>>
\mathbf\Sigma_E@>>>\Omega\times T\bC^2\\
&&@VV\pi_{\omega,\bZ_2}V@VV\pi_{\bZ_2}V\\
&&\Sigma_{\omega,E}@>\iota_{\omega,E}>>\Sigma_E
@>\pi_{{\mathcal L}}>>\hat\Sigma_E\phantom{\times T\bC^2}
\end{CD}
\end{equation*}
\caption{Overview over the relations between the different phase spaces}
\label{abb:orientierung}
\end{figure}
\begin{theorem}\label{thm:erg:Coul}
Let the shift action $\vartheta\colon{\mathcal L}\times\Omega\to\Omega$
be ergodic with respect to~$\beta$,
the potential $\tilde V_\omega$ three times continuous differentiable and
$(\tilde V_\omega,E)$ of strictly negative curvature for
$\beta$-almost all $\omega\in\Omega$.\\
Then the flow~$\hat\Phi_E$ is ergodic with respect to~$\hat\mu_E$.
\end{theorem}
\begin{proof}
In \cite{Hop41} the ergodicity of a recurrent and hyperbolic geodesic flow
on a $2$-dimensional riemannian surface is proven. We use that work here.
We need to show that for every continuous function
$\hat f\colon\hat\Sigma_E\to\bR$ the limits
\begin{equation}\label{fbar}
\bar f^\pm:=\lim_{T\to\pm\infty}\frac1T
\int_0^T\hat f\circ\hat\Phi_E(t,\cdot)\,dt
\end{equation}
exist and are almost everywhere constant.
The existence a.e.\ of~$\bar f^\pm$
is guaranteed by Birkhoff's ergodic theorem,
and so is $\bar f^+=\bar f^-=:\bar f$ $\hat\mu_E$-a.e..
Via the quotients and embeddings, see \cref{abb:orientierung}, we introduce
\begin{equation*}
f_\omega^*\colon\mathbf\Sigma_{\omega,E}^*\to\bR\textq,
f_\omega^*:=\hat f
\circ\pi_{\mathcal L}
\circ\iota_{\omega,E}
\circ\pi_{\omega,\bZ_2}
\circ\mathbf\Pi_\omega^*
\quad(\omega\in\Omega)\text.
\end{equation*}
By Poincar\'e's recurrence theorem
the finite measure preserving dynamical system
$(\hat\Sigma_E,\hat\Phi_E,\hat\mu_E)$ is recurrent, so that
$\liminf_{t\to\infty}\abs{f_\omega^*\circ(\mathbf\Phi_{\omega,E}^*)_t
-f_\omega^*}=0$ a.e.,
see \emph{e.g.} \cite[p.~14--16]{Aar97}.
Hyperbolicity is guaranteed by the curvature assumption.
Using the a.e.\ constancy of $\bar f_\omega^*$
along stable and unstable manifolds like in \cite{Hop41}
we see, that for each $\omega\in\Omega$ the limit~$\bar f_\omega^*$
is almost surely constant.
Finally we use the ergodicity of $(\vartheta,\beta)$
to conclude that $\bar f$ ist constant on~$\hat\Sigma_E$.
By denoting this a.e.-value by $F(\omega)$
for $\beta$-a.a.\ $\omega\in\Omega$
we define $\beta$-a.e.\ a function $F\colon\Omega\to\bR$.
Since the whole construction depends measurably on $\omega\in\Omega$,
$F$~itself is measurable.
To see that $F$ is in fact $\vartheta$--invariant, we introduce
\begin{equation}\label{fhat}
f_{\Sigma_E}:=\bar f\circ\pi_\c
\colon\Sigma_E\to\bR
\qquad(\omega\in\Omega)
\end{equation}
and write for a.e.\ $\omega\in\Omega$,
a.e.\ $x\in\Sigma_{\vartheta_\ell\omega}$ and a.e.\ $y\in\Sigma_\omega$
\begin{align*}
F(\vartheta_\ell\omega)&
=f_{\Sigma_E}\circ\iota_{\vartheta_\ell\omega}(x)
=f_{\Sigma_E}\circ\Theta^E\circ\iota_\omega(y)
=f_{\Sigma_E}\circ\iota_\omega(y)
=F(\omega)\text,
\end{align*}
using that $f_{\Sigma_E}$~by definition is $\Theta^E$--invariant,
see \cref{fbar,fhat}.
Therefore, by ergodicity of $(\vartheta,\beta)$,
$F$~is $\beta$-a.s.\ constant,
which then implies that $\bar f$~is constant modulo~$\hat\mu_E$.
\end{proof}
The above theorem relates to a geodesic flow on
a (extended and compactified) unit tangent bundle.
This construction has been made in order to make use of the
well-known consequences of negative curvature on the dynamics.\\
Like in the case of smooth potentials treated in \cref{sec5},
we are finally interested in the (analog of)
the Hamiltonian flow \eqref{tPhiE} generated by $\hat{H}$ (based on the Hamiltonian $H\colon P\to \bR$ on the extended phase space~$P$
of the first regularization in \cref{sec9}).
As this flow is related to the geodesic flow of \cref{thm:erg:Coul}
by a continuous time change, we get:
\begin{corollary}\label{cor:erg:Coul}
Under the conditions of \cref{thm:erg:Coul}
the coulombic Hamiltonian flow \eqref{tPhiE}
on the compactified energy surface $\hat{\Sigma}_E$
is ergodic w.r.t.\ Liouville measure $\tlambda_E$.
\end{corollary}
\begin{remark}[Markov Partition]\label{todo}
In a forthcoming article, one of us (CS) shows
for all large energies, under some additional conditions,
the existence of Markov partitions for the system, cf.~\cite{Sch10}.
\end{remark}
In \cite{CLS10} Cristadoro, Lenci and Seri show recurrence for
random Lorenz tubes. Their configuration spaces are contained in a connected
union of translates of a fundamental polygon.
We a have similar statement for the Coulomb case.
Again with $H\colon P\to \bR$ from \cref{sec9},
for $\ell\in{\mathcal L}\setminus \{0\}$ we define the factors
\[\check\Sigma_E:=H^{-1}(E)/\mathrm{span}_{\bZ}(\ell)
\qtextq{and}
\check\Sigma_{E,\omega}:=H_\omega^{-1}(E)/\mathrm{span}_{\bZ}(\ell)\text,\]
for $\ell$--periodic $\omega\in\Omega$.
\nomenclature[AXhat]{$\check\Sigma_{E,\omega}$}{AXhat...}{}{}%
\nomenclature[AXhatomega]{$\check\Sigma_{E,\omega}$}{Coulomb tube}{}{}%
Again we assume $(V_\omega,E)$ to be of strictly negative curvature $\beta$--a.s..
As we divided by the free $\Theta$--action of a subgroup of~${\mathcal L}$,
by the analog of \eqref{B} we obtain smooth flows on the non--compact
three--manifolds~$\check\Sigma_{E,\omega}$.
We assume for simplicity that $\ell\in{\mathcal L}$ is primitive.
Then ${\mathcal L}/\mathrm{span}_{\bZ}(\ell)\cong \bZ$
(and via the element $\ell'$ of a positively oriented basis
$(\ell,\ell')$ of ${\mathcal L}$, this isomorphism is unique).
We may consider random potentials indexed by $\Omega_\ell:=J^\bZ$.
\nomenclature[GOmegaell]{$\Omega_\ell$}{single site potential configurations on $\check\Sigma_E$}{}{}%
$\bZ$--ergodic probability measures $\beta_\ell$
\nomenclature[Gbetal]{$\beta_\ell$}{probability measure on $\Omega_\ell$}{}{}%
on $\Omega_\ell$ give rise to ${\mathcal L}$--ergodic image measures~$\beta$
on $\Omega$, via the injection $\mathbf{i}\colon\Omega_\ell\to\Omega$,
${\bf i}(\omega)_{a\ell+b\ell'}=\omega_b$.
\nomenclature[Ai]{$\mathbf i$}{injection $\Omega_\ell\to\Omega$}{}{}%
These are analogues of the measures considered in \cite{CLS10}.
For non--trivial $\beta_\ell$ the flow on~$\check\Sigma_E$ is not ergodic
w.r.t.\ the measure induced by~$\check\mu_E$
via the projection $\check\Sigma_E\to \hat\Sigma_E$.
\begin{proposition}[Ergodicity and Recurrence of Coulomb Tubes]
Under\\ the conditions of \cref{thm:erg:Coul}
\begin{compactitem}
\item $\beta$--a.s.\ the motion on~$\check\Sigma_E$ is recurrent.
\item For ${\bf i}(\beta_\ell)$--almost all $\omega\in \Omega$
the motion on~$\check\Sigma_{E,\omega}$ is recurrent and ergodic.
\end{compactitem}
\end{proposition}
\begin{proof}
By \cref{cor:erg:Coul} the flow~$\hat\Phi_E$
is ergodic w.r.t.\ $\tlambda_E$. Thus by the analog of
\cref{prop:ConsequencesOfErgodicity} for Coulomb potentials,
for $\beta$--a.e.~$\omega$
the asymptotic velocity ${\overline{v}}(x)=0$ a.e.\ on~$\Sigma_{E,\omega}$,
and thus on $\check\Sigma_E$.
If $\beta$ is of the form ${\bf i}(\beta_\ell)$, then for $\beta$--a.e.~$\omega$
the asymptotic velocity ${\overline{v}}(x)=0$ a.e.\ on~$\check\Sigma_{E,\omega}$.
We compare with Schmidt \cite{Sch98} (and the references cited therein).
In the case of an ergodic transformation~$T$ on
a standard probability space $(X,\mu)$, a Borel map $f\colon X\ar\bR$
and the induced cocycle (orbit sum) $\hat f\colon\bZ\times X\ar\bR$, he defines
$\sigma_k(A):=\mu(\{x\in X\mid\hat f(k,x)/k\in A\})\quad(k\in\bN)$,
for Borel sets $A\subseteq \bR$.
Under the assumption that vaguely $\lim_{k\ar\infty}\sigma_k=\delta_0$,
he notes that $f$~is \emph{recurrent}, meaning that
$\liminf_{k\ar\infty} \abs{\hat f(k,x)}=0$, $\mu$--a.s..
Here we use a Poincar\'e map discretization $(X,\mu,T)$
of the flow $\hat\Phi_E$ on~$\hat\Sigma_E$
(by \cref{todo} such a discretization exists and is ergodic).
For $f(x_0)$, with $x_0=[(\omega, p_0,q_0)]\in X$, we use the difference
$\LA q(T(x_0))-q_0,\ell^\perp\RA$ in position along the direction
$\ell^\perp:=\bsm 0&1\\-1&0\esm\ell\in \bR^2$ perpendicular to $\ell$.
$f$ is well-defined on~$X$.
The assumptions of \cite{Sch98} apply and $f$~is recurrent.
This means that for a.e.\ $x_0$
the motion in the Coulomb tube returns infinitely often to every neighbourhood
of the circle in the configuration torus given by $\LA q_0,\ell^\perp\RA=0$.
Then one obtains, by Poincar\'e's Recurrence Theorem for the induced map,
recurrence in the usual sense.
Ergodicity of the flow on~$\check\Sigma_{E,\omega}$ follows from local ergodicity,
using Hopf's argument \cite{Hop39}.
\end{proof}
\printnomenclature
\addcontentsline{toc}{section}{References}
\bibliographystyle{alpha}
|
\section{Introduction}
Horn and Weinstein\cite{HW84} and Horn et al\cite{HKW85} proposed the $t
--expansion for the calculation of the ground--state energy of
quantum--mechanical models. It is the Taylor expansion about $t=0$ of a
monotonically decreasing function $E(t)$ that leads to the ground--state
energy when $t\rightarrow \infty $ provided that the chosen reference
function exhibits a nonzero overlap with the ground state. The coefficients
of such Maclaurin series are known as connected moments or cummulants.
Since the extrapolation of the $t$--expansion towards $t\rightarrow \infty $
by means of Pad\'{e} approximants\cite{HW84,HKW85} did not appear to produce
encouraging results, Cioslowski\cite{C87a} proposed an exponential series
and Stubbins\cite{S88} compared it with other extrapolation approaches. On
matching the $t$--expansion and the exponential series at origin one has to
solve a system of nonlinear equations. In order to bypass this problem
Cioslowski\cite{C87a} developed a systematic algorithm for obtaining just
the parameter related to the energy by removing all the other variables in
the nonlinear equations. The resulting approach is known as connected
moments expansion (CMX). From the properties of the Pad\'{e} approximants
Knowles\cite{K87} derived a compact an ellegant explicit expression for the
approximants to the energy in terms of the connected moments.
Since the CMX approximants may exhibit
singularities\cite{K87,MBM89,MPM91} other authors proposed
improved approaches like the alternate moments expansion
(AMX)\cite{MZM94, MZMMP94} and the generalized moments
expansion(GMX)\cite {MMFB05}. Although these methods may avoid the
singularities in the CMX approximants they do not seem to improve
the convergence properties of the sequences of approximants and
commonly their results are not more accurate than the CMX ones.
A recent analysis of the convergence properties of the CMX
required the calculation of all the variables in the CMX
ansatz\cite{AFR11a}. For that reason, Amore and
Fern\'{a}ndez\cite{AF11b} proposed a systematic method for the
full solution of the CMX nonlinear equations motivated by an
earlier discussion\cite{F09} of the connected--moments polynomial
approach\cite{B08}.
Nonlinear equations like the CMX ones are an old problem in
applied mathematics and were first solved by Prony\cite{P1795} and
later Weiss and McDonough\cite{WM63} proposed an alternative
approach based on Pad\'{e} approximants. In fact, Prony's method
is well known in numerical analysis\cite{H74} and has been widely
applied to a variety of problems in chemistry and
engineering\cite{FMT07,HP02}. Further inspection of the procedure
developed by Amore and Fern\'{a}ndez\cite{AF11b} has revealed that
it is closely related to Prony's method, although the main result
derived by the former authors does not appear in the discussions
of the latter method\cite {WM63,H74}.
The main purpose of this paper is to show the connection between
Prony's method and the procedure developed by Amore and
Fern\'{a}ndez. In Sec.~\ref {Sec:Prony} we show how to solve a
particular set of nonlinear equations by means of Prony's method
and derive the main result of Amore and Fern\'{a}ndez\cite{AF11b}.
In Sec.~\ref{Sec:Generating} we discuss the application of that
method to the extrapolation of the generating functions for the
moments and connected moments. In Sec.~\ref{Sec:examples} we test
the general analytical results on simple models and show that they
are useful for a discussion of the convergence properties of the
CMX as well as for the approximate calculation of the
autocorrelation function.
\section{Prony's method}
\label{Sec:Prony}
In what follows we discuss the solution of the system of $2N$ equations
\begin{equation}
F_{k}=\sum_{n=1}^{N}A_{n}b_{n}^{k+s},\;k=1,2,\ldots ,2N \label{eq:main_eqs}
\end{equation}
for the $2N$ unknowns $A_{n}$ and $b_{n}$, where $F_{k}$ are known real
numbers and $s$ an integer.
Prony's method is based on the construction of the polynomial\cite{WM63,H74}
\begin{equation}
p(b)=\prod_{n=1}^{N}\left( b-b_{n}\right) =\sum_{j=0}^{N}p_{j}b^{j},\;p_{N}=1
\label{eq:poly}
\end{equation}
where the polynomial roots $b_{n}$, and thereby the polynomial coefficients
p_{j}$, are unknown. It follows from
\begin{equation}
\sum_{j=0}^{N}F_{i+j}p_{j}=\sum_{n=1}^{N}A_{n}b_{n}^{i+s
\sum_{j=0}^{N}p_{j}b_{n}^{j}=0,\;i=1,2,\ldots ,N
\end{equation}
that the coefficients $p_{j}$ are solutions to the linear system of
equations
\begin{equation}
\sum_{j=0}^{N-1}F_{i+j}p_{j}=-F_{i+N},\;i=1,2,\ldots ,N \label{eq:linear_pj}
\end{equation}
If the matrix $\mathbf{F}$ with elements $F_{i+j}$, $i=1,2,\ldots ,N$,
j=0,1,\ldots ,N-1$ is nonsingular then the solution is unique. Once we have
the polynomial coefficients $p_{j}$ we obtain its roots $b_{n}$ and then
solve the resulting system of linear equations (\ref{eq:main_eqs}) (for
example for $k=1,2,\ldots ,N$) for the remaining unknowns $A_{n}$.
The technique just outlined is known since 1795\cite{P1795} and is commonly
called Prony's method\cite{WM63,H74}. Weiss and McDonough\cite{WM63}
proposed an alternative way of solving the nonlinear equations by means of a
partial fraction expansion of Pad\'{e} approximants, and in what follows we
describe a different strategy for obtaining the roots $b_{n}$ developed by
Amore and Fern\'{a}ndez\cite{AF11b}. The starting point of this procedure is
the homogeneous system of $N$ linear equations
\begin{equation}
\sum_{i=1}^{N}\left( F_{i+j}-bF_{i+j-1}\right) c_{i}=0,\;j=1,2,\ldots ,N
\label{eq:secular_eqs}
\end{equation}
with $N$ unknowns $c_{i}$. There will be nontrivial solutions ($c_{i}\neq 0
) only if the determinant vanishes:
\begin{equation}
\left| F_{i+j}-bF_{i+j-1}\right| _{i,j=1}^{N}=0 \label{eq:secular_det}
\end{equation}
Under such conditions we define
\begin{equation}
\gamma _{j}=\sum_{i=1}^{N}F_{i+j}c_{i}
\end{equation}
so that equations (\ref{eq:secular_eqs}) become $\gamma _{j}=b\gamma _{j-1}$
that leads to $\gamma _{j}=b^{j}\gamma _{0}$, $j=1,2,\ldots ,N$. Therefore,
it follows from
\begin{equation}
\sum_{j=0}^{N}\gamma
_{j}p_{j}=\sum_{i=1}^{N}c_{i}\sum_{j=0}^{N}F_{i+j}p_{j}=\gamma
_{0}\sum_{j=0}^{N}p_{j}b^{j}
\end{equation}
that the roots of the determinant (\ref{eq:secular_det}) are exactly the
roots of the polynomial $p(b)$ of Prony's method.
We have now at least three alternative procedures for solving the nonlinear
equations (\ref{eq:main_eqs}). First, the original Prony's method that
consists of solving the system of linear equations (\ref{eq:linear_pj}) for
the coefficients of the polynomial $p(b)$, then calculating its roots $b_{n}
, and finally solving the resulting system of linear equations (\ref
{eq:main_eqs}) for the remaining unknowns $A_{n}$. Second, the partial
fraction expansion of the Pad\'{e} approximants proposed by Weiss and
McDonough\cite{WM63}. Third, our approach that consists of obtaining the
polynomial $p(b)$ from the determinant~(\ref{eq:secular_det}). The choice of
either of them is probably a matter of taste. We find our technique quite
straightforward and it has revealed a most interesting connection between
the exponential expansion of the generating function for the moments and the
Rayleigh--Ritz method in the Krylov space\cite{AF11b}.
Prony's method was developed for fitting a series of exponential functions
to given numerical data\cite{P1795,WM63,H74,FMT07}. In what follows we apply
it to a completely different problem so that the name Prony's method refers
only to the strategy for solving the nonlinear equations (\ref{eq:main_eqs}).
\section{Generating functions for the moments and connected moments}
\label{Sec:Generating}
The general equations developed in the preceding section prove useful for
the analysis of the CMX\cite{C87a,K87}. We first consider the
moments--generating function\cite{HW84}
\begin{equation}
Z(t)=\left\langle \phi \right| e^{-t\hat{H}}\left| \phi \right\rangle
\label{eq:Z(t)}
\end{equation}
where $\hat{H}$ is the Hamiltonian operator of the system and $\left| \phi
\right\rangle $ is an arbitrary reference state. The coefficients of its
Maclaurin series
\begin{equation}
Z(t)=\sum_{j=0}^{\infty }\frac{(-t)^{j}}{j!}\mu _{j} \label{eq:Z(t)_Taylor}
\end{equation}
are the moments $\;\mu _{j}=\left\langle \phi \right| \hat{H}^{j}\left| \phi
\right\rangle $.
For simplicity we assume that the spectrum of $\hat{H}$ is discrete
\begin{equation}
\hat{H}\left| \psi _{j}\right\rangle =E_{j}\left| \psi _{j}\right\rangle
,\;j=0,1,\ldots
\end{equation}
where $E_{0}\leq E_{1}\leq E_{2}\leq \ldots $, and without loss of
generality we choose the eigenvectors of $\hat{H}$ to be orthonormal:
\left\langle \psi _{i}\right| \left. \psi _{j}\right\rangle =\delta _{ij}$.
Under such conditions the generating function exhibits the exponential
behaviour
\begin{equation}
Z(t)=\sum_{j=0}^{\infty }\left| \left\langle \phi \right| \left. \psi
_{j}\right\rangle \right| ^{2}e^{-tE_{j}} \label{eq:Z(t)_exp_exp}
\end{equation}
where the expansion coefficients $\left| \left\langle \phi \right| \left.
\psi _{j}\right\rangle \right| ^{2}$and the energies $E_{j}$ are unknown. We
can calculate them approximately by means of the method developed in the
preceding section and the ansatz
\begin{equation}
Z_{N}(t)=\sum_{j=0}^{N-1}A_{j}e^{-tW_{j}} \label{eq:Z_N(t)}
\end{equation}
where $A_{j}$ and $W_{j}$ are the approximations to $\left| \left\langle
\phi \right| \left. \psi _{j}\right\rangle \right| ^{2}$and $E_{j}$,
respectively. If we require that the Maclaurin expansion of this ansatz
yields the first moments $\mu _{j}$ exactly we are left with the nonlinear
equations (\ref{eq:main_eqs}) with $F_{k}=\mu _{k-1}$, $s=-1$ and
W_{j}=b_{j-1}$.
It is has been proved that matching those expansions at origin is equivalent
to the application of the Rayleigh--Ritz variational method in the Krylov
space (RRK)\cite{AF11b}. Note that the determinant (\ref{eq:secular_det})
becomes the secular determinant of the RRK. The RRK solutions
\begin{equation}
\left| \varphi _{j}\right\rangle =\sum_{i=0}^{N-1}c_{ij}\left| \phi
_{i}\right\rangle ,\;\left| \phi _{i}\right\rangle =\hat{H}^{i}\left| \phi
\right\rangle \label{eq:varphi_j}
\end{equation}
satisfy
\begin{equation}
\left\langle \phi _{k}\right| \hat{H}\left| \varphi _{j}\right\rangle
=W_{j}\left\langle \phi _{k}\right| \left. \varphi _{j}\right\rangle
\label{eq:RRK}
\end{equation}
and the application of Prony's method in the way just outlined yields
A_{j}=\left| \left\langle \phi \right| \left. \varphi _{j}\right\rangle
\right| ^{2}$.
The CMX is based on the generating function
\begin{equation}
E(t)=-\frac{Z^{\prime }(t)}{Z(t)} \label{eq:E(t)}
\end{equation}
that is monotonically decreasing\cite{HW84} and the coefficients of its
Maclaurin series are the connected moments $I_{j}$:
\begin{equation}
E(t)=\sum_{j=0}^{\infty }\frac{(-t)^{j}}{j!}I_{j+1} \label{eq:t_exp}
\end{equation}
The recurrence relation\cite{HW84}
\begin{eqnarray}
I_{1} &=&\mu _{1} \nonumber \\
I_{j+1} &=&\mu _{j+1}-\sum_{i=0}^{j-1}\left(
\begin{array}{c}
j \\
i
\end{array}
\right) I_{i+1}\mu _{j-i},\;j=1,2,\ldots \label{eq:I_j(muj)}
\end{eqnarray}
yields the connected moments (or cummulants) $I_{i}$ in terms of the moments
$\mu _{j}$.
The main interest in $E(t)$ is that it provides a size consistent approach
to the ground--state energy\cite{HW84}
\begin{equation}
\lim_{t\rightarrow \infty }E(t)=E_{0}
\end{equation}
when $\left\langle \phi \right| \left. \psi _{0}\right\rangle \neq 0$. In
order to carry out this extrapolation Cioslowski\cite{C87a} proposed the
exponential--series ansatz
\begin{equation}
E^{(N)}(t)=A_{0}+\sum_{j=1}^{N}A_{j}e^{-b_{j}t} \label{eq:E^(N)(t)}
\end{equation}
where the unknown parameters $b_{j}$ are supposed to be real and positive
and $A_{0}$ is the approximation to $E_{0}$. Matching this expression with
the $t$--expansion (\ref{eq:t_exp}) leads to the set of equations
\begin{equation}
A_{0}=I_{1}-\sum_{n=1}^{N}A_{n} \label{eq:A0_I1-An}
\end{equation}
and
\begin{equation}
I_{k+1}=\sum_{n=1}^{N}A_{n}b_{n}^{k},\;k=1,2,\ldots ,2N
\label{eq:I_j_A_j_b_j}
\end{equation}
We can solve the set of $2N$ equations (\ref{eq:I_j_A_j_b_j}) by means of
the method of the preceding section with $F_{k}=I_{k+1}$ and $s=0$. In this
case the exponential parameters $b_{j}$, $j=1,2,\ldots ,N$ are the roots of
the pseudo--secular determinant
\begin{equation}
\left| I_{i+j+1}-bI_{i+j}\right| _{i,j=1}^{N}=0 \label{eq:secular_I_ij}
\end{equation}
Once we have them we solve $N$ of the resulting $2N$ linear equations (\ref
{eq:I_j_A_j_b_j}) for the coefficients $A_{j}$, $j=1,2,\ldots ,N$, and then
we obtain $A_{0}$ from equation\ (\ref{eq:A0_I1-An}). Cioslowski\cite{C87a}
and Knowles\cite{K87} developed remarkable strategies for obtaining $A_{0}$
without calculating the other parameters explicitly. We have just shown that
the explicit calculation of all the parameters in the exponential--series
ansatz (\ref{eq:E^(N)(t)}) is quite straightforward.
\section{Illustrative examples}
\label{Sec:examples}
In order to test the equations developed in the preceding section
and show their usefulness in the analysis of the CMX we first
consider an exactly solvable model: the dimensionless harmonic
oscillator
\begin{equation}
\hat{H}=-\frac{d^{2}}{dx^{2}}+x^{2} \label{eq:H_HO}
\end{equation}
As a trial function we choose one of the examples discussed in earlier paper
\cite{AF11b,AFR11a}
\begin{equation}
\phi (x)=\left\langle x\right| \left. \phi \right\rangle =\left( x^{2}-\frac
1}{2}\right) e^{-2x^{2}/5} \label{eq:phi(x)_HO}
\end{equation}
for which $\left| \left\langle \phi \right| \left. \psi
_{2}\right\rangle \right| >\left| \left\langle \phi \right| \left.
\psi _{j}\right\rangle \right| $, $j\neq 2$. (The normalization
factor is irrelevant to present purposes) One advantage of this
example is that it is not difficult to obtain the generating
function exactly
\begin{equation}
E(t)=\frac{121u^{3}+189199u^{2}+8180919u+6561}{\left( 81-u\right) \left(
121u^{2}+20198u+81\right) },\;u=e^{-4t} \label{eq:E(t)_HO_exact}
\end{equation}
Note that $\lim_{t\rightarrow \infty }E(t)=E_{0}=1$ because $\left\langle
\phi \right| \left. \psi _{0}\right\rangle \neq 0$ in agreement with the
discussion in Sec.~\ref{Sec:Generating}.
By means of the general results of the preceding section we can
easily calculate the generating function $E(t)$ approximately in
two ways: as $U^{(N)}(t)=-Z_{N}^{\prime }(t)/Z_{N}(t)$ from
equation (\ref{eq:Z_N(t)}) and directly from equation
(\ref{eq:E^(N)(t)}). Fig~\ref{Fig:ETHO2A} shows that $U^{(N)}(t)$
approaches the exact generating function $E(t)$ for all $0\leq
t<\infty $ as $N$ increases. On the other hand, the approximate
expression $E^{(N)}(t)$ given by Eq.~(\ref{eq:E^(N)(t)}) is an
unsatisfactory approximation to the exact $E(t)$ as shown in
Fig.~\ref{Fig:ETHO2B} (except for sufficiently small values of
$t$). This situation does not appear to improve as $N$ increases.
It is found that $\lim_{t\rightarrow \infty }E^{(N)}(t)=-\infty $
for $N=2,3$ because in both cases there is one negative root
$b_{j}$ associated to a negative coefficient $A_{j}$. For example,
$A_{1}\approx -0.0170$, $A_{2}\approx 0.147$, $A_{3}\approx
0.000617 $, $b_{1}\approx -3.87$, $b_{2}\approx 4.04$, and
$b_{3}\approx 9.29 $ for $N=3$. Note that the curves in both
figures are directly comparable in the sense that the calculation
of either $Z_{N+1}(t)$ and $E^{(N)}(t)$ requires exactly the same
number of moments $\mu _{j}$ ($2N+1$). The behaviour of
$E^{(N)}(t)$ for larger $N$ may be different; for example, for
$N=4,5$ there are two negative roots and $\lim_{t\rightarrow
\infty }E^{(N)}(t)=+\infty $.
The anomalous behaviour of $E^{(N)}(t)$ is due to the poles of
$E(t)$ in the complex $t$--plane. Note that
Eq.~(\ref{eq:E(t)_HO_exact}) exhibits three real poles in the
$u$--plane and the closest to the origin is located at $u\approx
-0.0040$. It is reasonable to assume that the CMX ansatz
$E^{(N)}(t) $ should approach the Taylor series about $u=0$ for
$E(t)$ when the approach is successful. However, this expansion
does not converge for $u=1$ which is necessary for matching the
$t$--expansion and exponential series at $t=0$. This problem was
discussed by Amore et al\cite{AFR11a} by means of two--level
models and we see it here again for the harmonic oscillator. Those
authors have shown that the CMX approximants for the harmonic
oscillator with the trial function (\ref{eq:phi(x)_HO}) converge
towards the second--excited state $E_{2}=5$. From the approximate
expressions $E^{(N)}(t) $ with $N=1,2,3$ we obtain $A_{0}\approx
4.932,5.015,5.002$, respectively, in agreement with those results.
This example shows that the CMX approximants may converge to a
meaningful result even when the ansatz $E^{(N)}(t)$ (on which the
approach is based) is an unacceptable global approximation to
$E(t)$.
Knowles\cite{K87} argued that if the connected moments $I_{k}$ are
positive for all $k$ then the $b_{i}$ are real and positive.
Later, Massano et al\cite {MBM89} and Mancini et al\cite{MPM91}
suggested that Knowles' argument may not be valid and that it is
the Hadamard determinants constructed from the $I_{k}$ that should
be positive. The example just analysed supports the latter
conclusion because the first connected moments are all positive
($\approx 5.13$, $0.665$, $2.12$, $11.2$, $40.0$, $216$, $979$)
but one of the roots is negative as shown above. We realize that
the general formulas derived in Sec.~\ref{Sec:Prony} are most
useful for this kind of analysis. In fact, the calculation of the
roots $b_{j}$ by means of present formulas is almost as easy as
the calculation of the Hadamard determinants and the former
provide a much clearer indication of the suitability of the ansatz
$E^{(N)}(t)$ as an approach for $E(t)$.
By means of the example discussed above we do not want to convey
the wrong impression that the CMX is suitable for the calculation
of excited--state energies because it is quite difficult to choose
an appropriate trial function for this purpose (except when we can
exploit the symmetry of the problem to make the trial function
orthogonal to the ground--state\cite {AF09b}). The main interest
in that exactly solvable problem is to show that we may predict an
anomalous behaviour of the CMX by means of the roots $b_{j} $. If
they are not real and positive it is reasonable to suspect that
the trial function may not be suitable and that it is convenient
to choose another one. For example, if $\phi (x)=e^{-x^{2}}$ all
the roots $b_{j}$ are positive and the resulting $E^{(N)}(t)$ is a
remarkably good approximation to $E(t)$ for all $0\leq t<\infty $
even at the low orders $N=1,2$.
Finally, we explore the possibility of calculating of the
correlation function
\begin{equation}
C(t)=Z(it)=\left\langle \phi \right| e^{-it\hat{H}}\left| \phi \right\rangle
\label{eq:C(t)}
\end{equation}
where $\left| \phi \right\rangle =\left| \phi (0)\right\rangle $
is the initial state normalized to unity and $\left| \phi
(t)\right\rangle =e^{-it\hat{H}}\left| \phi (0)\right\rangle $ is
the state at time $t$. In particular we concentrate on the real
function $\left| C(t)\right| ^{2}$.
As a first example we consider the harmonic oscillator (\ref{eq:H_HO}) and
the trial function
\begin{equation}
\phi (x)=\left( \frac{2}{\pi }\right) ^{1/4}\mathbf{e}^{-x^{2}}
\label{eq:phi(0)_HO}
\end{equation}
The exact correlation function is
\begin{equation}
\left| C(t)\right| ^{2}=\frac{4\sqrt{2}}{\sqrt{41-9\cos {\left( 4t\right) }}}
\label{eq:C(t)^2_HO}
\end{equation}
Fig.~\ref{Fig:ZITHO} shows results for $N=2,3,4,5$. There is no
doubt that $\left| Z_{N}(it)\right| ^{2}$ converges towards
$\left| C(t)\right| ^{2}$ as $N$ increases. The approximation is
satisfactory within the first period but the errors accumulate as
$t$ increases. This fact is not surprising because the approximate
expression is constructed from the Maclaurin series for $Z(it)$.
The second example is the anharmonic oscillator
\begin{equation}
\hat{H}=-\frac{d^{2}}{dx^{2}}+x^{4} \label{eq:H_AHO}
\end{equation}
and the same trial function (\ref{eq:phi(0)_HO}).
Fig.~\ref{Fig:ZITAHO} shows that the approximate results for
$N=2,3,4,5$ follow the same trend as in the case of the harmonic
oscillator. At first sight it may be surprising that the
convergence rate for the anharmonic oscillator appears to be
greater than for the harmonic one. The reason is that the chosen
trial state exhibits a greater overlap with the ground state of
the former. Note that the magnitude of the coefficient $A_{0}$ of
$Z_{N}(t)$ tells us that $\left| \left\langle \phi \right. \left|
\psi _{0}\right\rangle \right| ^{2}$ is approximately 0.943 for
the harmonic oscillator and 0.981 for the anharmonic one.
As a two--dimensional example we consider the anharmonic oscillator
\begin{equation}
\hat{H}=-\frac{d^{2}}{dx^{2}}-\frac{d^{2}}{dy^{2}}+x^{2}+y^{2}+\lambda
x^{2}y^{2} \label{eq:H_PE}
\end{equation}
where $\lambda >0$, and the trial function
\begin{equation}
\phi (x,y)=\left( \frac{2}{\pi }\right) ^{1/2}e^{-x^{2}-y^{2}}
\label{eq:phi(0)_PE}
\end{equation}
Fig.~\ref{Fig:ZITPE} shows results for $\lambda =0.5$ that look quite
similar to those for the one--dimensional models.
One does not expect that it may be possible to construct $C(t)$
from its Taylor expansion about $t=0$ in all the cases. The
problems discussed above are simple examples of single--well
oscillators. In the case of the double--well oscillator
$V(x)=(x^{2}-1)^{2}$ and the trial state localized in one of the
wells $\phi (x)=\left( \frac{2}{\pi }\right) ^{1/4}e^{-(x-1)^{2}}$
the approximants to $\left| C(t)\right| ^{2}$ do not appear to
converge although the results for $Z(t)$ look reasonable. If, on
the other hand, we locate the initial state on top of the barier
(for example, Eq.~(\ref{eq:phi(0)_HO})) then the results are as
accurate as those for the single wells.
\section{Conclusions}
\label{sec:conclusions}
The main goal of this paper is to show that\ Prony's method provides the
full solution of the CMX equations in a straightforward and simple way. We
have outlined the connection between Prony's method and the procedure
developed recently by Amore and Fern\'{a}ndez\cite{AF11b}. Although both
approaches are equivalent and Prony's one is known since long ago it seems
that the pseudo--secular determinants obtained by Amore and Fern\'{a}ndez
are not available elsewhere.
In spite of the fact that the harmonic oscillator with the trial function
\ref{eq:phi(x)_HO}) was chosen as an illustrative example in two earlier
papers\cite{AFR11a,AF11b} we think that present discussion based on the
exact generating function (\ref{eq:E(t)_HO_exact}) is clearer and provides
more information about the behaviour of the CMX ansatz $E^{(N)}(t)$. In
particular, we have shown that this simple problem clearly illustrates that
Knowles' conclusion about the sign of the connected moments\cite{K87} is not
valid as argued before by Massano et al\cite{MBM89} and Mancini et al\cite
{MPM91}.
It is clear that the roots $b_{i}$ provide a clear indication of the success
of the CMX and the method developed by Amore and Fern\'{a}ndez\cite{AF11b},
as well as Prony's method, make their calculation straightforward. Both
procedures may be applied to the calculation of the autocorrelation function
for simple quantum--mechanical models. If the approach converges then one
obtains a reasonable analytic expression for the autocorrelation function
within the first period of oscillation.
|
\section*{\abstractname}%
\else
\small
\begin{center}%
{\bfseries \ackname\vspace{-.5em}\vspace{\z@}}%
\end{center}%
\quotation
\fi}
{\if@twocolumn\else\endquotation\fi}
\fi
\makeatother
\begin{document}
\begin{center}
\textbf{
{\Large{Enhancing Reasoning Skills in the Process of Teaching and
Learning of Physics via Dynamic
Problem Solving Strategies: a Preparation for Future Learning}}\\
\medskip
{\Large{
Sergio Rojas \\
Departamento de F\'{\i}sica, Universidad Sim\'on Bol\'{\i}var \\ Venezuela.
}}
}
\end{center}
\begin{abstract}
The large number of published articles
in physics journals under the title ``Comments on $\cdots$''
and ``Reply to $\cdots$'' is indicative that the conceptual
understanding of physical phenomena is very elusive and hard to grasp
even to experts, but it has not stopped the development of Physics. In fact,
from the history of the development of Physics one
quickly becomes aware that, regardless of the state of conceptual
understanding, without quantitative reasoning Physics would
have not reached the state of development it has today.
Correspondingly, quantitative reasoning and problem solving skills are
a desirable outcomes
from the process of teaching and learning of physics.
Thus,
supported by results from published research,
we will show
evidence that a well structured problem solving strategy taught as a dynamical
process offers a feasible way for students to learn physics quantitatively
and conceptually, while helping them to reach
the state of an \emph{Adaptive Expert} highly
skillful on innovation and efficiency, a desired outcome from the perspective
of a \emph{Preparation for Future Learning} approach
of the process of teaching and learning Physics effectively.
\end{abstract}
\section{Introduction}
From the perspective
of a \emph{Preparation for Future Learning} approach
of the process of teaching and learning physics effectively,
a contemporary view
encourage
addressing the process so that students become
\emph{adaptive experts} \cite{SchwartzEtAll:2005,HatanoInagaki:1986},
who are individuals highly efficient in applying (transferring) what they know
to tackle new situations and are also extremely capable of innovation
in the sense of being able to inhibit inadequate
blocking
``off the top of the head processes''
or ``to break free of well-learned routines'' so that they can move to
new learning episodes by finding, perhaps ingenious,
ways to approach first time situations.
In this regard, it is undeniable that
\emph{Physics Education Research} (PER) and psychological research
on learning and instruction have
made available a good deal of teaching strategies
which are helpful in reaching the aforementioned goal
(a few such strategies are
listed in \cite{HendersonDancy:2009}
and some of them have been reviewed elsewhere
\cite{DocktorAndMestre:xxxx,LitzingerEtAl:2011,AmbroseEtAl:2010}).
Nevertheless, some controversial debates
in relation to the effectiveness of some of these \emph{Research-Based
Instructional Strategies} (RBIS) can also be found
in the literature
\cite{Sobel:2009a,Lasry:2009a,Sobel:2009b,Lasry:2009b,Glazek:2008,Klein:2007,KleinDebate1:2007,KleinDebate2:2007,KleinDebate3:2007}.
Nobel Prize winner Professor Carl Wieman has also called for
cautiousness when measuring RBIS teaching outcomes as one could create
illusions about what students actually
learn\cite{Wieman:2007a}. As mentioned in an article about transfer:
``standard methods of investigating transfer have
tended to depend on success-or-failure measures of participants' behavior
on transfer
tasks designed with very specific performance expectations on the
part of the investigator. These methods have failed to identify the
productive knowledge that
students often do bring to bear on the tasks given to them.''\cite{Wagner:2006}
In this regard, Professor Sobel in more emphatic
``Yes, in a special (possibly
grant{}-supported) program, with smaller groups, with highly motivated
instructors and students, with less content, students might do well,
but that's not the real
world.''\cite{Sobel:2009b}
Or as warned by Professors Reif and Allen ``Because nominal expertise does not
necessarily imply good performance, one must be cautious in interpreting
cognitive studies of novices and experts for which 'experts' have been chosen
on the basis of nominal criteria. Data about such experts must be interpreted
cautiously to avoid misleading conclusions about thought processes leading
to good performance.''\cite{ReifAllen:1992}
In this panorama,
of particular concern considering the intrinsically quantitative
nature of
physics is the fact that
in physics classes
students should actually be trained to apply what they have learned
in their math classes
and the aforementioned opinions might be a result of the fact that
many of the published papers on teaching and learning
physics
seem to overemphasize the
importance of teaching conceptual physical
aspects
\cite{HoellwarthMoelter:2011,MualemEylon:2007,SabellaRedish:2007,Walsh:2007,HoellwarthEtAll:2005},
and to deemphasize the significance of standard mathematical
reasoning\cite{Taber:2009}, which are crucial for understanding
physical processes, and
which are not stressed, or even taught, because, rephrasing a passage
from a recent editorial,
they interfere with the students' emerging sense of physical
insight. \cite{Klein:2007} A view which is further
stressed in a physics textbook instructor manual
(\cite{Knight:2008}, page 1-9):
``the author believes that for students struggling to grasp many
new and difficult concepts, too high a level of mathematics detracts
from, rather than aids, the \emph{physics} we want them to learn.
There is ample time in upper division courses for a more formal and
rigorous treatment. It's counterproductive to burden students with
unfamiliar and frightening mathematical baggage during their first
exposure to the subject.'' And these kind of opinions seems to be reflected
in outcomes obtained from the application of the relatively recently
developed CLASS survey, which measures student's
beliefs about physics and learning physics, showing a decrease of roughly
$15\%$ (out of 397) after instruction on student's beliefs about problem
solving in
physics
and a decrease of $12\%$ (out of 41) after instruction
on student's beliefs about the connection
between physics and mathematics
(see respectively tables I and V of \cite{AdamsEtAl:2006}. Both courses were
calculus-based Physics I. In relation
to table I (N=397), the authors of the study mention that
``These are typical results for a first semester course----regardless
of whether
it is a traditional lecture-based course or a course with interactive
engagement in which the instructor does not attend to student's
attitudes and beliefs about physics."
Moreover, controversial outcomes coming from some highly publicized
RBIS
\cite{Ates:2007,Coletta:2008,Ates:2008,KostSmithEtAl:2010}
hardly help physics instructors in
finding suitable advice about how to approach the
teaching of physics in the most efficient way and an answer to the
question of how much time should be spent on intuitive, conceptual
reasoning and how much time in developing quantitative reasoning.
In view of the aforementioned facts, the aim of this paper is to
begin a discussion on how
the process of teaching and learning
physics via \emph{dynamic problem solving strategies} can help
to tackle
not only the conceptual but also the quantitative reasoning
deficiencies
persistently
reported in the literature regarding the performance of students
in introductory and upper-division physics courses
\cite{WallaceChasteen:2010,Thompson:2009,Meredith:2008,MarshallCarrejo:2008,BoudreauxEtal:2008,Shaffer:2005,ReifAllen:1992,McDermott:1992:p1,Weeren:1982,LarkinEtAl:1980}.
It is obvious that both conceptual and quantitative reasoning
are desired skills students should acquire and develop in our
physics courses in order
to foster in them their willingness to explore more complex scientific
or engineering problems with confidence: \emph{a preparation
for future learning}.
The rest of the paper is organized as follows.
Recalling events
from the history of physics,
in the next section,
\emph{Conceptual versus quantitative understanding},
we argue that
in spite of conceptual gaps in some key physical ideas
the development of physics did not stop
because of the quantitative nature of physics. We also present
in this section an empirical result showing how a student changes his/her
wrong initial (conceptual) intuition
by using
quantitative
analysis when solving a problem about electrical circuits.
The next section discusses some of the needs for teaching students
the use and application of structural
problem solving strategies.
The following section, before presenting
the conclusions,
develops the central theme in the article:
promoting deep approaches to learning via \emph{Dynamic problem
solving strategies}. An illustrative example is also discussed
in this section.
\section{Conceptual versus quantitative understanding}
From a practical point of view, the conceptual understanding of the
principles of physics is a difficult and in some cases a very elusive
task. This is confirmed by the large body of research dealing with
``Student ideas about $\cdots$'',
``Student understanding of $\cdots$'', ``Students misconceptions about
$\cdots$'', ``students' misunderstanding of $\cdots$'' and so
forth.
For instance, in a
study \cite{HrepicEtAll:2007}
performed in a rather highly suitable and exceptionally favorable
teaching environment, it was found that students answered correctly
some conceptual questions
on the nature of sound propagation in the same proportion
before and after
receiving instruction on the subject. In that study,
in addition to active teaching instruction, students also
watched at their own pace video-lectures on the nature of
sound propagation
by an experienced instructor, Professor Paul Hewitt.
After the
pre-instruction test, the students
were told
that each one of the test questions will be
answered in these video-lectures.
Similar results on students performance have also been reported
from a study about why
the seasons change.
Students maintained their conceptual misconceptions about the subject
even after watching a video that clearly explained the phenomena
\cite{AmbroseEtAl:2010}.
Perhaps more dramatic is the repeatedly reported case in which
students respond to conceptual questions about the behavior of physical
quantities the same way as students did at the beginning of
the 80's \cite{Shaffer:2005}.
Should we be surprised
by these findings? Not, at all. As mentioned by Ambrose and collaborators
``It is important to recognize that conceptual change occurs gradually
and may not be immediately visible. Thus, students may be moving in the
direction of more accurate knowledge even when it is not yet apparent
in their performance'' \cite{AmbroseEtAl:2010}.
In fact, this is also observed in well trained physicists.
For example, great debates about
the proper understanding of the concept of physics have taken place
among brilliant physicists. To be specific,
in this regard
one could refer to
the conceptual debates in statistical and quantum
physics and the electromagnetic
theory \cite{Biro:2011,Plotnitsky:2010,BragaEtAll:2010,BacciagaluppiValentini:2009,Lindley:2001,Caneva:1980,Whittaker:1910};
the controversies
between Lorentz and Einstein on the conceptual understanding and the
meaning of the principles
of special relativity \cite{Janssen:2002};
and in other areas \cite{Lange:2011,Bunge:1966}.
What is even more relevant to PER is the fact that
many of these controversies persist
still today among expert physicists who invest a great deal of their time
thinking about and working with these matters
\cite{Swendsen:2011,Hobson:2009,Kampen:2008,Baierlein:2006,Lalo:2001,JacksonReplyRoche:2000,RocheReplyJackson:2000,Jackson:1999}.
Additionally, the large number of ``Comment on $\cdots$'' and
``Reply to $\cdots$'' articles in physics journals are also a reminder
of the difficulty
of understanding and applying the concepts of physics.
Nevertheless, in spite of the aforementioned controversies on the
conceptual understanding
of physical principles, the development of physics has not stopped. One could
argue that the reason for it
is rooted in the fact that
``nature is too subtle to be described from any single point of view.
To obtain an adequate description, you have to look at things from
several point of views, even though the different viewpoints are
incompatible and can not be viewed simultaneously.''\cite{Dyson:1991}
And,
as a matter of fact,
physics is fortunately a combination of two basic compatible viewpoints
``physical reasoning''
and ``quantitative reasoning''.
Correspondingly,
in spite of debates on the nature
and significance of the concepts of physics,
the intrinsic quantitative nature of physics is what has
propelled the development of physics,
helped by the
sometimes questioned
\emph{Scientific Method} \cite{Chalmers:1999,Giunta:2001}.
While the scientific method
helps us to organize and test systematically
every single hypothesis (enhancing our conceptual understanding)
\cite{Kipnis:2010,Bassow:1991},
quantitative reasoning helps us to be precise
in which body of knowledge (mathematical models of the physical world)
needs to be further developed:
those \emph{de accord} or that are consistent with
observations and experiments \cite{Chalmers:1999,Peierls:1991,Peierls:1979}.
To make the point clearer,
one could think about the kind of progress physics
would have reached had not Kepler
struggled to fit the orbit of Mars to an elliptical one,
stopping because of his lack of understanding (provided by Newton
around 80 years later) of why the orbits of the planets were
following the laws he was uncovering. Or think
about the course of knowledge had Galileo given up his view
of doing experiments and finding mathematical explanations
for them
in favor of the Catholic Inquisition's conceptual ideas about the universe.
Or think about the current state of development in physics had Planck (because
of lacking the respective conceptual understanding) restrained himself
from introducing (in 1900) the Planck's constant to
resolve the ultraviolet
catastrophe. Or think about Einstein not continuing the
development of his Theory of General
Relativity because of the lack of conceptual understanding
for not keeping in his theory
the cosmological constant leading to a static universe.
But, have expert physicists today overcome the conceptual understanding
undermining
those developments? The answer in no.
To be explicit in one case, one could see how physicists are still
trying to understand quantum mechanics at its deepest conceptual level,
a problem which arose more than one hundred years ago,
at the beginning of the last century, with
the introduction of Planck's constant. Yet, in spite of the conceptual
shortcoming, the mathematical formulation of quantum mechanics and
its refinements have allowed
physics, regardless of the conceptual gap,
to progress to levels that in today's world at many
physics and engineering
research centers,
researchers are making conclusive observations
about the nano-scale world for unanimated matter.
Thus, in each one the aforementioned cases, and of the many others that
can be cited,
there is no doubt that it was the quantitative analysis
undertaken by the scientist
involved that raised further the value
of scientific knowledge, even when at the time it was hard to provide
satisfactory conceptual explanations of the phenomena (such explanations
came much later, after further developments of the mathematical understanding
of each phenomena).
Without them one would be
talking today about ``philosophical or scriptural proclamations'' rather
than scientific ones.
Consequently, each one of these facts speak about the necessity of
having our students
of science and engineering
to become properly acquainted with quantitative reasoning in their early
training, even if they
are lacking deep conceptual understanding. In this way they will be able
to further deepen their understanding as they arrive to study upper
division courses.
\begin{figure*}[htb]
\begin{center}
\leavevmode
\includegraphics[height=.2\textheight]{circuit}
\end{center}
\caption{The circuit contains an ideal battery, three identical light bulbs
and a switch. Initially the switch is open. After the switch closes, does
the brightness of bulb A increase? Explain.
This question was given
as a post-test written examination
to a total of 23 students at the University of Maine this year
(2011). Only two students
attempted to answer the question quantitatively. The others
applied rather unsuccessfully a \emph{Case Based Reasoning}
approach \cite{Kolodner:1992}, trying to answer the question recalling
from memory some classification patterns devised for conceptual
understanding
of this type of problems \cite{McDermott:1992:p2}.
}
\label{fig:circuit}
\label{page:Figcircuit}
\end{figure*}
In terms of the teaching and learning of physics,
an empirical example of
how quantitative reasoning can help students to accurately
reason conceptually, even though the students' initial intuition
might be wrong, can be seen in data
from a post-test written examination given to students
enrolled in a first year introductory physics course at the University
of Maine (UMaine). The formulation of the problem is shown in
Figure \ref{fig:circuit}. At the start,
answering the question intuitively, one student
wrote that the brightness of bulb A should decrease.
Then the student went on to explain why it would happen that way as follows:
``The brightness should decrease
because the brightness of the bulb depends on the current of the circuit.
So when the switch is open to find the current of the two light bulbs in series
you would use the formula
$I_{\text{total}} = V_{\text{total}}/R_{\text{total}}$. So let say $V=12 v$
and each bulb acts as a resistor with $2\Omega$ of impedance. When it's
open there are two resistors in the circuits so $I = 12/(2+2)=3 A$
compared to when it's closed we simplify it to a series circuits so
the $R_{\text{total}}$ of the parallel would be $1 \Omega$ so
$1 \Omega + 2 \Omega = 3 \Omega$ as $R_{\text{total}}$. $V=12 v$
so $12/3 = 4 A$ so when the switch is thrown there's more current and the
bulb is brighter''.
Analysis of this and other students' answers of this study will
be published elsewhere \cite{Hawkins:2011}.
Here
we should only notice
how the use of quantitative reasoning helped the student to
correct his/her initial (wrong) intuition
to the correct result that
after closing the switch bulb A becomes brighter.
Thus, this example provides
evidence that the learning
with
emphasis on equations and stressing
the conceptual physical meaning of the respective
symbols in the equation is a very feasible task. In particular,
proper guidance and additional training in applying physical equations
with understanding
will
help this student to reason analytically, using symbols, instead of
resorting to numerical
values (though we are not against this practice, specially in more
difficult situations), which is more useful
when dealing with situations on which physical intuition might
fail \cite{Singh:2002} (assuming that resistance of bulb $A$ is
known, a non-intuitive question
regarding the circuit of figure \ref{page:Figcircuit}
would be to ask about values for the resistance of bulbs $B$ and $C$
in
order to have a maximum power in that section of the circuit
when closing the switch. This is a kind of
open-ended problem which
has multiple solutions.)
From the cognitive
point of view, the helpfulness of quantitative reasoning in conceptual
understanding can be grounded in the assumption
that
``...the use of mathematics in physics presupposes measurements. Measurements
transform difficult to relate perceptual quantities into a common numerical
ontology that supports precise comparisons and
relations.''\cite{SchwartzEtAll:MathPhys:2005}
Thus, while ``guiding students through a process of conceptual change
is likely to take time, patience, and creativity''\cite{AmbroseEtAl:2010},
students needs to becomes acquainted in applying what they have been
learning quantitatively, so they can readily use that knowledge
appropriately in
their more advanced courses. To make this possibility a reality
one needs to devote time to approaching the process of teaching and
learning physics with emphasis on equations and the meaning, not only
of the equation but also of each one of the symbols in the equation
(for example, the meaning and consequences of
$y=y_0 + m x$
when $m$ represents a constant acceleration are different from the
situation on which it ($m$) might represents
a constant speed. In each case,
the symbols, $y$, $y_0$ and $x$, might have different meanings too. For
instance, in the case of $m$ representing constant acceleration, $x$
could represent either position, then $y$ and $y_0$ should
represent the square of a velocity, or time, then $y$ and $y_0$ should
represent a speed).
In this way, one could minimize PER findings on students being
proficient in manipulating formulas without understanding
the meaning of the symbols in the equation \cite{Sherin:2001}.
Correspondingly, as pointed out
by Professor Hewitt \cite{Hewitt:2011}
``Isn't teaching emphasis on symbols
and their meanings in an introductory [physics] course a
worthwhile effort?''
\section{On why a problem solving strategy is needed}
Another worrying outcome from the learning and instruction of physics
and science research literature
is the observed lack of ability by students
to apply a structured reasoning
methodology that could, among other things,
help them to identify the nature of a problem as well as
the principles and
quantitative models
to solve it.
An example can be found in a study \cite{Walsh:2007}
reporting that
out of 22 students solving a set of six physics problems, 9
``Analyzes the situation based on required variables.
Proceeds by choosing formulas based on the variables in a trial and error
manner.'', 6 ``Proceeds by trying to use the
variables in a random way.'', 2 ``Proceeds by trying to 'fit' the
given variables to those
examples.'', and 5 ``Plans and carries out solution
in a systematic manner based
on that analysis.'' Similar observations can be drawn from the
analysis of interview excerpts
reported in other studies \cite{BoudreauxEtal:2008,Wagner:2006,Sherin:2001}.
This lack of a structural way of reasoning is also
observed in the responses to the circuit
question presented in
the previous section
(see Figure \ref{fig:circuit}). In relation to the same question,
another
student reasoned as follows:
``No, Decrease. Adding more resistance drop the current. Using the \#'s I
used, power decreases $\therefore$ Brightness goes Down." Writing the
answer the student wrote
$\text{Brightness} = P = IR^2$ (she/he wrote the following
numbers next to every circuit element
$V=6 v$ for the battery and $2\Omega$ at each bulb). The student
continued with $A+B = 4 \Omega$ and $A + (1/B + 1/C) = 3 \Omega$. Then she/he
wrote $V=I R$; $V/R = I \Longrightarrow 6v/4\Omega = 1.5 A$;
$V/R = I \Longrightarrow 6v/3\Omega = 2 A$; $P = I R^2 = 1.5\cdot(4)^2 =24W$
and $2\cdot(3)^2 =18W$.
In addition, several studies have shown how initial thought processes
block students'
thinking
preventing them from going beyond a circular way of
reasoning \cite{Kahneman:2002,BilalicEtAl:2008}.
To mention an example, in analyzing the interview of a student solving a physics
problem it is reported that ``Dee-Dee was very unwilling to give up her
qualitative ideas about force and motion, even though she has already
written down the correct algebraic form of Newton's second law. Her
qualitative and quantitative dynamics knowledge appear to be associated
(she articulated them very close together in time), but they have not
been reconciled into a consistent knowledge
structure.''\cite{SabellaRedish:2007}.
Thus, in all of the mentioned cases we can observe
in most of the students
the lack of a structured and systematic reasoning
strategy which could
guide them to further analyze, verify, and make sense of the solution
procedure they use when solving a problem.
Additionally, this fact has also been reported in many studies comparing novice
and expert reasoning
abilities \cite{Singh:2002,ReifAllen:1992,ReifHeller:1982,LarkinEtAl:1980}.
The most common observed behavior in students is the plugging of numbers
in equations without much hesitation.
As a matter of fact, in a study \cite{Ogilvie:2009}
on which students were asked
to write down self-reflections on problem-solving one of them wrote
``Instead of learning the material and then doing the
assignments, I would just try to search for equations in the
text that would solve the problem, and if I couldn't figure it
out, just guess.'' Other student wrote
``The way I approach physics problems is by looking for a
formula to follow. I will start a problem, look in my notes for
a formula, look on the discussion board for a formula, look
on the formula sheet for a formula, and lastly, look in the
book for a formula. This is probably exactly I will approach
some problems in Computer Science. If I need to implement
a function, I will need to find the syntax for the function and
a little excerpt saying what inputs it has and what outputs it
has.''
Unfortunately, that behavior is encouraged
in some instructional settings \cite{Hamed:2008} and
throughout the summary of equations at the end of each chapter on
commonly used textbooks. And it is further stressed by the way
illustrative examples are worked out in those textbooks
\cite{Hewitt:2011,Kamal:2011,Rojas:2010rmf}.
Correspondingly, the
problem solving methodologies
commonly found in many textbooks
have been heavily criticized in PER
literature \cite{KimPak:2002,DocktorAndMestre:xxxx}. And we agree with
such criticism because what is preached and mostly illustrated in
commonly recommended textbooks for introductory physics
is a mechanical and static problem solving methodology:
find an equation, plug in the numbers, and get the answer,
taking away the joy of finding and exploring the different
ways of solving a problem and the making sense not only of the solution
procedure but also of the obtained answer.
More disappointing,
textbooks
examples end when the answer to the problem is found.
At most, in some few cases, there is a checking for dimensional consistency
of the result,
and in very rare cases one can find a discussion about the
feasibility of the obtained result.
But in general,
there is no further exploration of the solution procedure
(i. e. whether or not it is a logically applicable to
the situation at hand) and
no advice is given to students
on how they can be certain that the followed solution procedure is correct
(as a matter of fact, many wrong solution procedures found in textbooks
have been
reported in the literature without being corrected
by textbook writers
\cite{Saccomandi:2010,Bohren:2009,HutzlerEtAl:2004,Sandin:1973}).
More importantly,
no
advise is given to students regarding the fact
that some incorrect solution procedures
can lead to a correct solution \cite{Rojas:2010rmf}
or that some solution procedures
can be applied or transfered to solve other problems in different contexts
(just to mention one example, the computation of the gravitational and
the electrical field of
a mass and a charge distribution respectively
share some similarities, but that is
not mentioned in most commonly used textbooks. Additional examples
on this matter are mentioned in \cite{Rojas:arxiv:2010}).
Correspondingly, students don't get trained
in developing a sense of solving problems by analogy. And they might
even think that the mathematics used in physics is different from
the mathematics
they learn in their calculus classes \cite{Rojas:2008}. Furthermore,
in most commonly used textbooks, physics
equations are not fully discussed to properly connect the physical or
conceptual meaning of each symbol with the place the symbol has
in the equation (i.e. why and what does it mean that the symbol is
multiplying? or why and what does it mean that it appear as a negative
exponential factor? What happens if the value of a symbol increases?,
and so on).
This lack of further analyzing the significance of the symbols in
equations might explain difficulty of students reasoning in
situations involving multivariable
equations \cite{Thompson:2009,Kuhn:2009,Kuhn:2008,diSessa:2008}.
Moreover, the inadequacy of text-book worked-out examples to enhance
students learning has been made
evident in a study by Chi and collaborators \cite{ChiEtAll:1989}.
In the study it was found that the textbook worked-out examples did not
provide any clue for students make generalizations of the underlying
domain theory, a necessary training in order to help students
to develop important skills to
successfully transfer and apply the acquired knowledge in more complex
contexts.
Accordingly, the need for explicitly incorporating on the teaching
of physics a well structured \emph{dynamic problem solving strategy}
(introduced and explained in the next section)
is thus justified.
As a matter of fact, in the same way as the \emph{Scientific Method}
can be used to decide
among competing theories, a \emph{dynamic problem solving strategy} is a
means by
which students can
guide their thoughts
in deciding on the plausibility or reasonability
of each one of the steps taken when
solving a (physics)
problem. Furthermore, a \emph{dynamic problem solving strategy} helps
students to make sense
of what they are learning by fitting it into what they already know or believe.
Via the questioning of intermediated
results and asking questions about what is being done,
students can
detect flawed/wrong conceptual and/or computational procedures.
As students gradually make reflections on what is being done while solving
a problem, they
start to create or strengthen
their own ``mental library'' of what works and what does not work.
As quoted by the Nobel Laureate in Economic Sciences (1978)
Herbert Simon (cited in \cite{AmbroseEtAl:2010})
``\emph{Learning results from what the student
does and think and only from what the student
does and think. The teacher can advance learning only by influencing
what the student does to learn}.''
Thus, introducing high school and university students to a
\emph{dynamic problem solving strategy} will enhance and strengthen
their argumentative thinking skills
and will also help them to organize their reasoning skills by focusing
their mental effort. It additionally could help to internalize early in
students the fact
that (i) the process for solving scientific problems
requires creative
thought, the use of available resources (books, computers, articles, etc.),
and personal interaction with peers and colleagues. (ii) There
may be no simple answer to questions that have been posed. In some
cases the outcome of a calculation can be contrary to what is expected
by physical intuition. In other instances an approximation that
appears feasible turns out to be unjustified, or one that looks
unreasonable turns out to be adequate.
(iii) There might be several competing ``correct'' answers based
on available knowledge and the careful and judicious application
of a \emph{dynamic problem solving strategy} is necessary to pick
the correct one.
And (iv) the full understanding
of a problem and its solution requires both the quantitative formulation
of the problem and the conceptual
meaning of the symbols that appear in the equation(s) describing
the problem quantitatively.
In short, the learning of a \emph{dynamic problem solving strategy} would
be the starting point to generate a culture of reasoning and of making sense
in the classroom, a necessary condition in
``preparation for future
learning''
\cite{Osborne:2010,Fortus:2009,SchwartzEtAll:2005,HatanoInagaki:1986}.
\section{A dynamic problem solving strategy}
An important issue to be resolved in PER is the lack of an integrative
theoretical framework that can make sense of the richness
of the large amount of empirical results
collected via the many
\emph{Research Based Instructional Strategies}
(RBIS) generated in the field (for a partial listing of RBIS see
\cite{HendersonDancy:2009}). In this regard, some obstacles
need to be overcome. In addition to controversial
outcomes
mentioned in the introduction concerning some RBIS,
important ongoing debates alluding to fundamental issues associated with
the psychology of learning physics
\cite{Slotta:2011,HammerEtAl:2011}
indicate that we are still far from reaching such a theoretical
framework. Until then, inspired by the scientific method framework,
we find it plausible to enhance the teaching and learning
of physics via \emph{dynamic problem solving strategies} in order to
strengthen students' quantitative problem solving skills,
a necessity which is further stressed by the occurrence of
some disasters which have been associated to computing
mistakes \cite{Feldman:2010,StrogatzEtAl:2005}.
In general terms, a \emph{dynamic problem solving strategy},
not to be confused with \emph{modes of reasoning} (see below),
could be constructed from the following steps \cite{Rojas:2010rmf}:
(1) understand and describe the problem;
(2) provide a qualitative description of the problem;
(3) plan a solution;
(4) carry out the plan;
(5) verify the internal consistency and coherence of the equations used
and the applied procedures; and
(6) check and evaluate the obtained solution.
One step more or one step less, the aforementioned
\emph{problem solving strategy} looks similar, you
might rightly wonder, to any other problem
solving strategy commonly found in textbooks and which have
been heavily criticized
in PER literature \cite{KimPak:2002}. In fact, as mentioned earlier
we agree with those
criticism. For one additional reason,
once introduced, the strategy is not
consistently applied in the textbook illustrative examples, much
less in the student and instructor companion manuals. For another,
the steps of the problem solving strategy are presented as rigid steps
which need to be followed in the particular order they are written
and applied without connection or interaction between them
(certainly such an inflexible problem solving strategy can only
be of limited value). And finally,
textbooks end of chapter problems are
constructed and organized in such a way that students only need to find the
right equation to plug-and-chug some numbers to get
the answer to the problem. Correspondingly, consequences
of such strategies are reflected
in the findings of some PER studies (\cite{KimPak:2002} and
references there in).
On the other hand, a study comparing a typical textbook problem
solving strategy
with a consistently and coherently applied
strategy reports on the overwhelming advantage of the latter
in relation to the former
\cite{HellerReif:1984}.
To be specific, the authors of the comparative study propose
a prescriptive
theoretical model of effective human problem solving
according to which the problem solving process embraces
three major stages:
1) generation of an initial problem description (including qualitative
analysis) which is helpful in the construction of a problem solution;
2) obtaining the solution using appropriated methods; and 3) evaluation
and improvement of the solution. General comments on the nature
and significance of each stage
are given in \cite{ReifHeller:1982}.
In the referred work \cite{HellerReif:1984}, Professors Heller and Reif
discussed further details of the controlled study they performed
comparing the quantitative problem solving performance of three groups:
one guided by their proposed problem solving strategy, the other
group guided by typical textbooks problem solving directions, and
a comparison group working without any external guidance.
In addition to showing the remarkable superiority of the
participants
trained according to the
authors methodology,
three major lessons can be drawn from the study:
a) that participants guided by a problem solving strategy
performed better than those who did not have any guidance; b) that
``completeness and explicitness of procedures for constructing initial
problem description'', missing from textbooks problem solving
strategies, are crucial for attaining better problem solving performance;
and c) that the participants implemented any systematic problem
solving procedure
fairly easily once they became familiar with it, but to get to that level,
the procedure needs to be applied constantly,
consistently and coherently because ``human subjects tend to be fallible
and distractable, prone to forget steps in a procedure or to disregard
available information.'' The authors comment that
after the resistance was overcome,
some participants remarked that the steps ``really
work'' and that the problems seemed suddenly ``easy'' to solve,
showing in turn that students beliefs could be changed via a well
designed and applied problem solving strategy, while
at the same time strengthening their computational skills
(similar results are reported in a study in mathematics \cite{Higgins:1997}.
The author mentions that ``Compared with the students who
had received traditional mathematics instruction, the students who had
received problem-solving instruction displayed greater perseverance in
solving problems, more positive attitudes about the usefulness of
mathematics, and more sophisticated definitions of
mathematical understanding.'')
At this point we now turn to the topic of this section trying to answer
the question:
what, then, makes a \emph{dynamic problem solving strategy}?
First,
the implementation
of the methodology is not an inflexible static process (i.e. there
is not a specific order
for implementing the considered steps and one can avoid some of the
steps or even add a new one). For instance,
working via inductive reasoning one could start from the answer
to an unknown problem
(i.e. an observed natural phenomena) and work backward to provide
a well posed problem whose solution is the observed answer. Students
can be trained in this process via well designed and worked out exercises by
presenting them with
a careful and detailed
backward
analysis of the followed
solution process. More importantly, verification of a solution procedure
via backward analysis,
in addition to
helping students
build confidence
in the obtained solution, also guides them to organize information in their
memory by making connections with
prior knowledge, a condition which increases the accessibility of useful
knowledge.
Thus, going back and forth in the application of a
\emph{dynamic problem solving strategy}
provides the required
feedback that experts apply when dealing with the solution
of new situations.
This flexibility helps
to avoid getting trapped or stuck by top-of-the-head thoughts,
preformed convictions, or intuitions that don't help in going
forward \cite{Kahneman:2002,BilalicEtAl:2008}.
One just continue
developing and applying (mentally and/or in writing)
thought processes
until a solution is found, further analyzed and reconciled with
intuitions. We can either change the wrong intuition (i.e. as
one student did when solving the exercise
of figure \ref{fig:circuit}) or strengthen
the correct one. Additional guidelines on what sort of analysis should
be included can be extracted from the work of Professors
Polya \cite{Polya:1945}, Schoenfeld \cite{Schoenfeld:1980,Schoenfeld:1992},
and Reif \cite{Reif:2008}.
This way of building understanding can be contrasted with the mechanical
way in which textbook's illustrative examples are
worked out.
Second,
proper application
of a
\emph{dynamic problem solving strategy}
requires the consistent use of
any applicable \emph{mode of reasoning}:
deductive and/or inductive reasoning; reasoning via analogy and/or via
counterexamples; reasoning by \emph{reductio ad absurdum};
and many others, including rule, case, model, and
collaborative-collective modes of reasoning.
These modes of reasoning
are rarely mentioned in commonly used introductory physics textbooks.
Nevertheless, in their mathematics courses
students might have already studied some of these modes of reasoning
(i.e. when proving by inductive reasoning the convergence of a sequence),
and it is
in their physics courses where they should have the opportunity to further
explore such modes of reasoning
from different perspectives, and became acquainted
with them.
Third,
a very important
skill in applying
a \emph{dynamic problem solving strategy} is the ability to ask questions.
Questioning, particularly at the higher cognitive
levels, is an essential aspect of problem
solving. By asking questions while solving a problem
one becomes engaged in a process of
self-explaining components of the underlying theory being applied
to solve the problem and
that were not explicitly exposed when learning the theory. Asking
questions also helps in the
detection of ``comprehension failures'' and in taking action to overcome
them. As put by Chi and collaborators ``Good students ask very
specific questions about what they don't understand. These specific
questions can potentially be resolved by engaging in self-explanations.''
\cite{ChiEtAll:1989}
Thus, at each step of a \emph{dynamic problem solving strategy}
one needs to stop and ask oneself about the significance
of what has been done so far, trying to find meaningful associations between
the new knowledge being applied and related concepts one already might know.
Examples of questions to be asked
constantly include:
how is this knew knowledge related to what I already know?;
in which context have I seen this problem before?;
in which context could I use this piece of knowledge?;
how are these seemingly disparate discrete pieces of knowledge
functionally and causally related?; can the principles to be
applied be used in this situation?; is this approach the right one?;
how can I be certain of it?; are you sure you can do that?;
how could this procedure be wrong?.
Asking questions helps students to be fully confident that each
step given while solving a problem is a reasonable one and do not
involve contradictory or implausible assumptions. Also, asking questions helps
to seek
evidence to contradict an accepted fact and it helps to strengthen confidence
in a solution procedure if we can set out different scenarios to investigate
the issues involved in a problem, as it requires engagement to identify
proper and reasonable ways to approach the problem and to think about
the interrelationships between the proposed approaches.
Fortunately, the education research literature has provided a good deal
of research on how one can help students develop the
habit of asking questions
\cite{GraesserPerson:1994,ZeeMinstrell:1997,HarperEtAll:2003,Chin:2010a,Chin:2010b}.
But, students will not
get the benefit of this process unless they
are explicitly taught how to use them. In this sense, being an intrinsic
part of a
\emph{dynamic problem solving strategy}, teaching it will develop in students
the habit of asking questions, a process
through which students could
build new useful knowledge that
they can then use in further developments as
they engage themselves in productive thinking and learning, not only
within the teaching and learning environment but also outside it.
We finalize this section in the hope of have made it clear that the
delivery of instruction following a \emph{dynamic problem solving
strategy} will create in students the habit of looking at problems
carefully and from multiple perspectives, choosing to be more mindful about
the making of sense of each solution procedure as they engage themselves
in the exploration of alternative ways of solving problems and of finding
explanations. The development of such ways of thinking certainly requires time,
effort, practice, self-reflection,
and feedback (from peers and the instructor). This is
how experts become experts, developing new intuition to cope properly
with situations where there is no right answer.
\subsection{Illustrative example}
In general, most problem solving strategies found in the literature
only mention checking the feasibility of the obtained final
result. Since some wrong procedures could lead to right results,
we have found it necessary to include an additional step
(step 5 mentioned at the beginning
of this section) in
constructing a \emph{dynamic problem solving strategy} so
faulty reasoning schemes could be detected \cite{Rojas:2010rmf}.
To illustrate the aforementioned thoughts, one could mention the
fallacious application of the idea of
\emph{separation of variables} (borrowed from solving partial differential
equations \cite{Arfken:1985})
to prove in two
dimensions the constant acceleration kinematic
relationships
$v_x^2 = v_{0x}^2 + 2 a_x(x-x_0)$
and
$v_y^2 = v_{0y}^2 + 2 a_y(y-y_0)$ (here
$\vec{\mathbf{a}}= a_x \hat{\mathbf{x}} + a_y \hat{\mathbf{y}}$,
$\vec{\mathbf{v}}= v_x \hat{\mathbf{x}} + v_y \hat{\mathbf{y}}$, and
$\vec{\mathbf{r}}= x \hat{\mathbf{x}} + y \hat{\mathbf{y}}$
represent respectively the acceleration, the velocity, and the displacement
of a particle, while
$\vec{\mathbf{v}}_0= v_{0x} \hat{\mathbf{x}} + v_{0y} \hat{\mathbf{y}}$, and
$\vec{\mathbf{r}}_0= x_0 \hat{\mathbf{x}} + y_0 \hat{\mathbf{y}}$ represents
respectively initial velocity and initial position of the particle.
$\hat{\mathbf{x}}$ and $\hat{\mathbf{y}}$ are orthogonal unit vectors).
Starting from $d\vec{\mathbf{v}} = \vec{\mathbf{a}} dt$, dot product
both sides of this equation by $\vec{\mathbf{v}}$ and rearrange terms
in the form
$d({v}^2/2) = \vec{\mathbf{a}}\mathbf{\cdot}\vec{\mathbf{v}}dt =
\vec{\mathbf{a}}\mathbf{\cdot}d\vec{\mathbf{r}}$, which after integration,
considering $\vec{\mathbf{a}}=\text{constant}$, yields
$v^2 = v_0^2 + 2\vec{\mathbf{a}}\mathbf{\cdot}(\vec{\mathbf{r}}
-\vec{\mathbf{r}}_0)$. This equation can be written in the form
\begin{equation}
v_x^2 - v_{0x}^2 - 2 a_x(x-x_0) = -(v_y^2 - v_{0y}^2 - 2 a_y(y-y_0)).
\label{eq:separada}
\end{equation}
So far, there is nothing wrong in any of the given steps
(in passing, let's mention that
by asking questions according to a \emph{dynamic problem solving strategy}
a student could gain conceptual understanding about the obtained
relationship: when is it applicable?, why is time dependence not explicit?
can it be applied to free fall?, can it be applied to projectile motion?
if so, under which conditions?, etc.)
The faulty reasoning starts when a student, wrongly invoking the
separation of variables technique, considers that
equation (\ref{eq:separada}) is in that form and, according to the
technique, the student
proceeds to
write that $v_x^2 - v_{0x}^2 - 2 a_x(x-x_0) = \alpha$ and
$v_y^2 - v_{0y}^2 - 2 a_y(y-y_0) = - \alpha$,
with $\alpha=\text{constant}$.
Now, considering that at $\vec{\mathbf{r}} = \vec{\mathbf{r}}_0$,
$\vec{\mathbf{v}} = \vec{\mathbf{v}}_0$, then $\alpha=0$ and
the proof is obtained, the student believes.
Thus, by means of following typical textbook
problem solving strategy, the student, without questioning any
of the solution steps, will consider that the proof is
correct because the
obtained answer is correct, and to show that that it
is so the student
could point to any introductory
physics textbook on which the same result is obtained by other procedures.
At this point one needs to mention that it is really hard to convince
students that despite leading to the correct answer, a solution procedure
could be mistakenly wrong. How, the student might wonder,
can it be that a wrong
solution procedure could yield the right response?. That does not make
any sense!!!
How can a \emph{dynamic problem solving strategy}
be helpful in making explicit the faulty reasoning?
The trick is on step 5 mentioned at the beginning
of this section:
once a solution procedure path has been established, one needs to
verify the consistency of each given
step (a missing step
from
any textbook problem solving strategy). And this can be attained
by working backward after the solution is obtained and
by asking questions at each step, wondering and justifying the
rightfulness of each one.
In this case, before calling
for the separation of variables technique, the student needs to ask about
whether the technique is applicable or not to the situation at hand. Since
the method of separation of variables
comes from a differential equation context,
a first question to ask, following a \emph{dynamic problem solving strategy},
is how to write equation (\ref{eq:separada}) so it looks like a differential
equation. This can be achieved by writing down the definition for each velocity
component in the equation, namely
$v_x = dx/dt $ and $v_y = dy/dt$. After doing so it turns out
that the working equation is actually a nonlinear first order differential
equation on which the time variable $t$ is common to both sides of
the equation, making
it impossible to apply the separation of variables procedure (which
requires that each side of the equation (\ref{eq:separada}) be a function
of only independent variables each to another. In this case both sides
of the equation
turn out to be dependent of the time $t$ variable).
Another way to settle the issue following a
\emph{dynamic problem solving strategy} is to wonder whether
a counter example can be built to rule out the use of the separation
of variables in this case. It turns out that by using basic calculus
a very simple
algebraic problem can be posed to see if the technique is or not applicable:
solve $(2x-x^2) = -(2y-y^2)$.
Applying the proposed separation of variables,
instead of an infinite set of solutions, only a finite set of solutions
will be found (namely
$x=0$ and $y=0$ or $y = 2$; $x=2$ and $y=0$ or $y = 2$).
Nevertheless, in spite of the evidence,
some students might still be reluctant to accept the wrongness
of their solution procedure and can start to
mention inexistent theoretical frameworks, like
a \emph{separation of coordinates} without
being able to provide
any reference to backup such a reasoning.
In this situation, only self-reflection and careful analysis of
the quantitative procedure can be used as a way to solve the dilemma.
Episodes like this are not difficult to find in PER
literature \cite{Sherin:2001},
and they can be described as ``reasoning to obtain the desired
result'', something that brings to memory what is
called Einstein's biggest mistake \cite{Weinberg:2005,Davies:1996},
which is associated with the
introduction by Einstein of a cosmic repulsion term
in order to model, in spite of the mathematical evidence, a static universe.
Could this be called Einstein's conceptual misunderstanding? or is it
simply a case in which
a good thought blocks a better one?\cite{Kahneman:2002,BilalicEtAl:2008}
\section{Concluding remarks}
Understanding that physics is essentially a quantitative based
subject,
in this article we have largely argue that the teaching
of solving problems in introductory physics courses
via a \emph{dynamic problem solving strategy}
can help students to develop quantitative reasoning
skills (necessary in the short term), while
enhancing their capabilities to develop conceptual understanding
via the analysis of the equation(s) representing
physical phenomena followed by the correct interpretation of the
physical meaning of each symbol in the respective equation(s)
\cite{GaigherEtAl:2007} (for instance, as mentioned earlier,
the meaning and consequences of
$y=y_0 + m x$
when $m$ represents a constant acceleration are different from the
situation on which it might represents
a constant speed. In each case,
the other symbols, $y_0$ and $x$, might have different meanings too. For
instance, in the case of $m$ representing constant acceleration, $x$
could represent either position, then $y$ and $y_0$ should
represent the square of a velocity, or time, then $y$ and $y_0$ should
represent a speed).
At each step of a \emph{dynamic problem solving strategies}
students ask questions that guide them to be fully confident that each
step given while solving a problem is a reasonable one and do not
involve contradictory or implausible assumptions.
Thus, this view of enhancing the teaching and learning of introductory
physics courses via \emph{dynamic problem solving strategies}
is de accord with the finding
that ``Good students
(those who have greater success at solving problems) tend to
study example-exercises in a text by
explaining and providing justifications for each action.
That is, their explanations
refine and expand the conditions of an action, explicate the
consequences of an action, provide a goal for a set of actions,
relate the consequences
of one action to another, and explain the meaning of a set of
quantitative expressions.''\cite{ChiEtAll:1989}
Our proposal is further
reinforced by the fact
that teaching directly or indirectly (i.e.
via scaffolding tutoring)
specific problem-solving strategies
improves students' scientific critical thinking and reasoning skills
\cite{Osborne:2010,MasonSingh:2010,Chi:2009,Nickerson:1988}.
As mentioned in a study
``When students used the procedural specification, they did
so properly and obtained correct answers - although they did not always
implemented all steps explicitly and resorted to
some shortcuts.''\cite{LabuddeReifQuinn:1988}
All of that is in agreement with the fact
that
learning to approach problems in a systematic way starts from
learning the interrelationships among conceptual knowledge,
mathematical
skills and logical reasoning.\cite{HeronMeltzer:2005}
And
it is in physics courses where
students can strengthen their quantitative reasoning
skills and even become acquainted with innovative non-standard
ways of solving problems
\cite{OlnessAndScalise:2011,BernsteinFriedman:2009}.
Thus, since the process of teaching and learning can rarely be done
in a completely closed and controlled environment, a well learned
\emph{dynamic problem solving strategy}
will equip students with reasoning capabilities
to properly question advice (easily found
on the
Internet and in
some articles \cite{Taber:2009})
that encourages them to memorize results and to
ignore the mathematical analysis leading to those results.
In other words,
considering that
``it is not possible to expect novice students to become more successful
problem solvers by simply telling them the principles which govern
the way experts sort physics problems." \cite{ChiEtAll:1989}
a well structured problem solving strategy taught as
a dynamical
process offers a feasible way for students to learn physics quantitatively
and conceptually, while helping them to reach
the state of an \emph{Adaptive Expert}, highly
skillful on innovation and efficiency: \emph{a preparation for future
learning}.
\begin{acknowledgements}
The author thanks the members of \emph{The Physics Education Research
Laboratory} (PERLab) at UMaine, USA, where this work
was done while on Sabbatical in the Department
of Physics and Astronomy.
\end{acknowledgements}
|
\section*{\sc Description}
We present a mission concept, the Geostationary Antenna for Disturbance-Free Laser
Interferometry (GADFLI), for a space-based gravitational-wave interferometer consisting of three satellites in geostationary orbit
around the Earth\footnote{After submission of this mission concept to the NASA-issued Request for Information (RFI), we became aware of a
very similar mission concept also submitted to the RFI \cite{Massimo}. That concept and the one we propose, though strikingly
similar, were nonetheless developed independently and submitted to the RFI simultaneously.}. Compared to the nominal design of the
Laser Interferometer Space Antenna (LISA), this concept has the advantage of significantly decreased requirements on the telescope size and
laser power, decreased launch mass,
and substantially improved shot noise resulting from the much shorter 73000 km armlengths ($2\times \cos(30^{\rm o})\times 42164$ km $\approx 73000$ km).
We note that the constellation and the size of the Earth are approximately to scale in the graphic on the title page.
Communications are also simplified in this design, and the possibility of a servicing mission in the event of a single satellite
failure is more viable. The three satellites are identical, and the cost of a 120 degree phase change for 2 satellites
is minimal, so that the principle launch cost will be the transfer of these three satellites, which are lighter than the three
LISA satellites, to a single geostationary orbit.
The primary disadvantage is the potentially diminished performance at frequencies below a mHz due to increased proof mass
acceleration noise.
A perceived secondary disadvantage is the need for station-keeping, since the Sun and Moon provide
torques out of the constellation's orbital plane due to its 23 degree inclination to the ecliptic, which would cause
a relative drift among the sciencecrafts. However, this drift is at the level of $\Delta v=45$ m/s each year,
and is primarily directed out of the plane of measurement, which implies a Doppler shift well below 45 MHz per year for a micrometer wavelength laser.
Conservatively estimating the Doppler shift at 25 MHz, phasemeter sampling at 50 MHz would therefore allow operation for at least two years
without station keeping.
Such a sampling rate is thought to be consistent with the capabilities
of available phasemeters in the range of cost of the nominal LISA design \cite{Ira}. We therefore limit the mission lifetime
of this proposal to 2 years, to avoid the need for the additional hardware required for station keeping and its impact on cost.
We consider three scenarios, where the noisier environment in Earth's orbit results in a failure to meet the DRS specifications of
$3 \times 10^{-15} {\rm m}/{\rm s}^2/\sqrt{\rm Hz}$ and performance is worse by an order of magnitude, where the DRS specification is met despite
the noisier environment, and the very optimistic scenario where the DRS specification is exceeded by an order of magnitude. Given that none
of these levels of accuracy has been achieved at these frequencies before, there is considerable uncertainty regarding what will be achievable
when the final mission hardware is assembled. We note that our optimistic choice places the residual DRS noise at the same level
as the galactic foreground of white dwarf binaries, so that further improvement would not benefit the science performance. This also
motivates our use of the term ``displacement free'' in the concept name, since an order-of-magnitude improvement over the nominal
LISA DRS would render displacement noise a non-factor in this design.
Compared to the OMEGA mission, which is another geocentric constellation with a much longer armlength,
GADFLI again has decreased requirements on telescope size and laser power, lower mass and decreased shot noise, and may have the same disadvantage
of increased acceleration noise, though this is less clear than in the comparison with LISA. An additional advantage of GADFLI is the 23 degree
inclination of an equatorial geocentric orbit with respect to the ecliptic, which will prevent the GADFLI optics from ever being exposed
to near-direct sunlight. This will also provide greater thermal stability, as the satellites are never eclipsed by the Earth over the course of their
orbit. Because of the longer armlength, OMEGA must orbit in the Earth-Moon plane, which is only inclined 5 degrees to the ecliptic, and must
therefore be outfitted with yet-to-be-developed filters, in order to prevent damage to the optics during their exposure to direct sunlight.
\section*{\sc Specifications}
The GADFLI constellation will fly at the $\sim$35800 km elevation above sea level
required for geostationarity, which corresponds to a constellation armlength
of approximately 73000 km. The mirror size can be decreased to 15 cm without resulting in a significant impact to sensitivity.
Likewise, the laser power can be reduced to 0.7 W. In Fig.~1, we plot the rms strain sensitivity
for three variations of the GADFLI concept, as well as the nominal LISA design and the recently-selected NGO configuration.
We have followed \cite{Larson} to implement the sensitivity, including the Hils and Bender white dwarf confusion
estimate \cite{Hils}. We have verified that, apart from the sinusoidal contributions to the response that we have neglected and which
are unimportant for signal-to-noise ratio (SNR) calculations, we duplicate the results found from running the online sensitivity curve generator \cite{Sens}
provided by the lead author of \cite{Larson}.
Although the noise ``bucket'' of the GADFLI curve is less sensitive, the frequency band of GADFLI is more optimal given the expectation of
likely sources, as we will show in the following section.
\begin{center}
\begin{figure}
\includegraphics[width=0.95\textwidth]{./sensitivities.ps}
\caption{Comparison of the rms sensitivity of several mission designs. The blue curve shows the nominal LISA design
with white dwarf binary confusion noise, the green curve shows the final NGO mission configuration, and the red, magenta, and black
curves show the concept described here, with 10, 1, and 0.1 times the acceleration noise permitted under the DRS specifications
for the nominal LISA design, respectively.
\label{fig:sens}
}
\end{figure}
\end{center}
\section*{\sc Science Performance}
Like LISA, GADFLI will be optimally sensitive to massive black-hole binary inspiral-merger-ringdowns occurring at the hearts of merged galaxies.
In Fig.~2, we plot contours of SNR for LISA and for the three variations of the proposed concept for these
signals. In addition,
we plot contours of the number of events for two merger-tree models, corresponding to small and large initial mass seeds \cite{Sesana}.
Given the superior sensitivity of GAD-HI compared to NGO, and the comparable sensitivity of GAD-MED (see Fig.~2), we expect both
of these designs to be sensitivity to extreme mass ratio inspirals (EMRIs) and individually resolvable galactic binaries, in addition to comparable
mass massive binaries. GAD-LO will likely be unable to observe EMRIs or galactic binaries, although it still has a significant
event rate for massive black-hole binary mergers from either merger catalog.
While we do not present detailed parameter estimation at this time, we expect the parameter accuracy to scale roughly inversely with SNR.
As with the nominal LISA design, masses and spins of merging black holes will be measured with extraordinary accuracy,
and any EMRIs that are detected will also provide accurate probes of spacetime, since the complex harmonic structure of
EMRI signals rapidly saturates the parameter measurement capability at a given SNR level.
It is noteworthy that, since the majority of the sky localization accuracy and spin measurement accuracy accumulates at the end of
the coalescence, the greater ease of communication with a geostationary constellation could assist in the earlier identification
of an electromagnetic counterpart to these events. If a counterpart illuminates during the final merger, a more frequent downlink
of data to the ground, combined with real-time data analysis, could either identify a gravitational wave source earlier, or
localize a known source on the sky earlier.
\begin{center}
\begin{figure}
\includegraphics[width=0.95\textwidth]{./snr_hists.eps}
\caption{Comparison of the signal-to-noise ratios (SNRs) among the nominal LISA design (blue), the $10\times$ DRS GADFLI (GAD-LO, red), the
$1\times$ DRS GADFLI (GAD-MED, magenta), and the $0.1\times$ DRS GADFLI (GAD-HI, black). In addition, we plot event number contours
for mergers from merger tree simulations consisting of small seeds (brown, $\langle \log(M)\rangle=3.7$)
and large seeds (green, $\langle \log(M)\rangle=5.3$). We note that the contours are all binned in a 30x30 logarithmically-spaced grid, so that
many bins, each containing between 1 and 10 events, lie inside the outermost number contours. The large seed catalog contains
a total of 720 sources (nearly all of which have SNR$>10$ with GAD-HI), and the small seed catalog contains a total of 2437 sources.
The SNR contours are calculated for an inspiral-merger-ringdown of massive black hole binaries with mass ratios of 3:1,
corresponding to the mean ratio in both catalogs.
\label{fig:hists}
}
\end{figure}
\end{center}
\section*{\sc Costing}
Given our lack of qualifications to perform a proper costing assessment, we will only provide a rough estimate, emphasizing that we expect
the lighter mass and smaller $\Delta$v orbit to be the principle areas of savings relative to the nominal LISA design. We follow the SGO
costing approach of subtracting cost from the \$1.8B SGO high price point (based on LISA with a less expensive launch vehicle). We assume
a modest mass savings relative to SGO lowest, so that the Falcon 9 (Block 2) from SpaceX would be the most cost effective
launch vehicle capable of supporting a launch mass slightly below SGO Lowest, for a cost savings of \$300M for savings
on the payload mass and launch vehicle costs. The 2 year mission lifetime provides a further savings of \$200M in personnel cost. Like
SGO lowest, GADFLI may not require a propulsion module, and certainly will need far less propellant than OMEGA or heliocentric designs.
We will assume the thrusters available or a less expensive propulsion module will be capable of executing the necessary
120 degree phase change for two sciencecraft that are already on the necessary geostationary orbital trajectories, which would provide
an additional savings of \$100M. The final cost estimate is therefore \$1.2B, though we again emphasize the rough nature of this estimate,
and the need for a much more thorough assessment by an actual costing expert.
\section*{\sc References}
|
\section{Introduction}
In this article, I consider orthogonal polynomials on
the unit circle, whose Verblunsky coefficients are given
by
\begin{equation}\label{eq:defVeromegank}
\alpha_n = \lambda \mathrm{e}^{2\pi\mathrm{i}\cdot \omega n^{k}}
\end{equation}
for $0\neq \lambda\in{\mathbb D}=\{z:\quad |z|<1\}$,
$\omega$ an irrational number, and $k\geq 2$.
The case $k=1$ corresponds to rotated versions
of the Geronimus polynomials, see Theorem~1.6.13
in \cite{opuc2} and Proposition~\ref{prop:rotatealpha}
(see also Theorem~5.3 in \cite{gzborg}).
Given Verblunsky coefficients $\alpha_n$,
we define orthogonal polynomials
on the unit circle recursively by
\begin{equation}\label{eq:defPhin}
\Phi_0(z) = 1, \quad \Phi_{n+1}(z) = z \Phi_n(z)
-\overline{\alpha_n} \Phi_n^{\ast}(z),
\end{equation}
where $\Phi_n^{\ast}(z) = z^n \overline{\Phi(\overline{z}^{-1})}$
is the reversed polynomial. By Verblunsky's theorem,
there exists an unique probability measure $\mu$ on $\partial{\mathbb D}$ such that
the $\Phi_n$ are orthogonal with respect to it.
The first result is
\begin{theorem}\label{thm:main1}
The support of $\mu$ satisfies
\begin{equation}
\mathrm{supp}(\mu) = \partial{\mathbb D}.
\end{equation}
\end{theorem}
The key to the proof of this theorem is that the
support of $\mu$ is the same as the support
of the measure with Verblunsky coefficients
$\alpha_n \mathrm{e}^{2\pi\mathrm{i} y n}$ by ergodicity for
any $y\in{\mathbb T}={\mathbb R}/{\mathbb Z}$. Now these two supports are
just rotated versions of each other. Hence
$\mathrm{supp}(\mu)$ must be the entire unit circle.
I give the details of the proof in Section~\ref{sec:pfmain1}.
Next, consider the family of Verblunsky coefficients
given by $\alpha_{x,n} = \alpha_n \cdot \mathrm{e}^{2\pi\mathrm{i} x}$.
The corresponding measures are known as {\em Aleksandrov
measures} $\mu_x$ see Section~3.2. in \cite{opuc1}.
Then we have that
\begin{theorem}\label{thm:main2}
For almost every $x$, the Aleksandrov measure $\mu_x$
is pure point.
\end{theorem}
The proof of this theorem is essentially the same as
Theorem~\ref{thm:main1}, since the rotational invariance implies
positivity of the Lyapunov exponent. Pure point spectrum
then follows from spectral averaging.
Deterministic examples with similar properties
have been previously obtained in \cite{dk}.
Adapting the methods of \cite{krho1}, \cite{krho2}
to orthogonal polynomials on the unit circle, it should
be possible to obtain similar even for $k > 1$
not an integer.
\bigskip
At this point, let me mention that the corresponding
question for orthogonal polynomials on the real line
respectively better Schr\"odinger operators is open.
Consider the potential $V(n) = \lambda \cos(2\pi \omega n^2)$
for an irrational number $\omega$. Then under a
Diophantine assumption on $\omega$ and a largeness
condition on $\lambda$ one can show pure point
spectrum, see \cite{b2002}, \cite{bgs},
and Chapter~15 in \cite{bbook} and that
the spectrum contains intervals \cite{kskew}.
However, it is believed that for all $\lambda > 0$
the spectrum of this operator is an interval
and pure point. Partial results for $\lambda > 0$
small can be found in \cite{b, b2,b3}.
The proofs of Theorem~\ref{thm:main1} and \ref{thm:main2}
are much easier than the real case, because of algebraic
miracles (Proposition~\ref{prop:rotatealphas}). However, there is also an analytic reason
why the case on the unit circle should be simpler,
namely that then the spectrum has no edges.
For this reason, I expect it to be possible to show
analogs of Theorem~\ref{thm:main1} and \ref{thm:main2}
if one perturbs $\alpha_n$ slightly by for example
$\alpha_n + \varepsilon f(\omega n^k)$ for an analytic
and one-periodic function $f$ and $\varepsilon > 0$
small enough.
\medskip
At first sight Theorem~\ref{thm:main1} and \ref{thm:main2}
might not seem too surprising, since we know many measures
whose support is the entire unit circle. But the Verblunsky
coefficients of these measures behave quite differently,
for regular measures one knows \cite{si} that the
Verblunsky coefficients Ces\'aro sum to $0$. Similarly
non-zero periodic potentials have at least one gap.
The situation becomes even more striking when considering
Schr\"odinger operators. There have been a series of
innovative works \cite{abd1,abd2,aj1,aj2,gs4,gsfest} to prove
Cantor spectrum, whereas there are only the perturbative
methods from \cite{cs,kskew} to prove that the spectrum
contains an interval.
\bigskip
\begin{figure}[ht]
\includegraphics[width=0.9\textwidth]{CMVGaps2N2000}
\caption{Zeros of $\Phi_{2000}(z; \beta)$ for $k=2$
and $\omega = \sqrt{2}$.}
\label{fig:1}
\end{figure}
Finally, I also want to address the zero distribution
of the para-orthogonal polynomials. This question has
not been discussed for Schr\"odinger operators yet.
Define for $\beta\in\partial{\mathbb D}$
\begin{equation}
\Phi_n(z; \beta) = z \Phi_{n-1}(z) - \overline{\beta} \Phi_{n-1}^{\ast}(z).
\end{equation}
In difference to $\Phi_n(z)$ the zeros of $\Phi_n(z; \beta)$
are on the unit circle. Denote these zeros
by $\mathrm{e}^{2\pi\mathrm{i} \theta_1}, \dots, \mathrm{e}^{2\pi\mathrm{i}\theta_N}$.
An inspection of the proof of Theorem~\ref{thm:evskew}
shows that an appropriate adaption of the
results would remain true for $\Phi_n(z)$.
Before stating our main result,
I will now illustrate the behavior of the zeros with
some numerical computations. Order the values
$\theta_j$ such that
\begin{equation}
0 \leq \theta_1 < \theta_2 < \dots < \theta_N < 1.
\end{equation}
Define the length of gaps by
\begin{equation}
g_j = \theta_{j+1} - \theta_j.
\end{equation}
Figure~\ref{fig:1} and \ref{fig:2} show the distribution
of the values of $g_j$ for different values of $N$
when $k = 2$. One
sees that this distribution peaks at only three values.
This should remind one of the distribution of gap
lengths for the sequence of values $\{\eta n \pmod{1} \}_{n=1}^{N}$
for some value of $\eta$ and in fact, we will show this
in Theorem~\ref{thm:main4}. Also it should be pointed
out that these gap distributions do not converge.
\begin{figure}[ht]
\includegraphics[width=0.9\textwidth]{CMVGaps2N4000}
\caption{Zeros of $\Phi_{4000}(z; \beta)$ for $k=2$
and $\omega = \sqrt{2}$.}
\label{fig:2}
\end{figure}
On the other hand Figure~\ref{fig:3} shows the same
graphic for $k =3$ and the distribution resembles
an exponential distribution. One obtains similar
figures for $k\geq 4$. This is the same distribution
one would obtain if the $\theta_j$ were given by
a Poisson process and by \cite{st06} also if the
the Verblunsky coefficients $\alpha_n$ were given
by independent identically distributed random variables
whose distribution is non constant and rotationally invariant.
\begin{figure}[ht]
\includegraphics[width=0.9\textwidth]{CMVGaps3N2000}
\caption{Zeros of $\Phi_{4000}(z; \beta)$ for $k=2$
and $\omega = \sqrt{2}$.}
\label{fig:3}
\end{figure}
Finally, in the case $k=1$, the (rotated) Geronimus Polynomials,
the assumptions of the Freud--Levin theorem hold
(Theorem~2.6.10 in \cite{szego}) and one has
clock spacing, so the spacing is given by the inverse
of the corresponding density of states measure.
This measure turns out to be non-constant, so
there is not a single peak.
In order to state our result, we need to
introduce more notation.
Define the Laplace functional of $N$ points
$x_1,\dots,x_N \in{\mathbb T}$ by
\begin{equation}
\mathfrak{L}_{\underline{x},N}(f) =
\int_{{\mathbb T}} \exp\left(-\sum_{n=1}^{N} f(N x_n(\theta))
\right)
d\theta,
\end{equation}
where $[-\frac{1}{2}, \frac{1}{2}) \ni
x_n(\theta) = x_n - \theta \pmod{1}$ and $f\geq 0$
is continuous and compactly supported function.
See \cite{kisto09} for a discussion of Laplace
functionals related to zeros of paraorthogonal
polynomials.
Denote by $\mathfrak{L}^{R}_{\omega, N}$
the Laplace functional of the sequence of points
$\{n\omega \pmod{1}\}_{n=1}^{N}$. The behavior of this sequence
is well understood, see for example \cite{raven}.
In particular, this quantity does not converge to
a limit. We will show
\begin{theorem}\label{thm:main4}
Let $k=2$, $\tau > 1$ and assume that
$\omega$ satisfies
\begin{equation}\label{eq:conddiop}
\inf_{q \geq 1, q\in{\mathbb Z}} q^{\tau} \dist(q \omega, {\mathbb Z}) > 0.
\end{equation}
Then for any positive, continuous, and compactly supported
function $f:{\mathbb R} \to{\mathbb R}$, we have
\begin{equation}
\lim_{N \to \infty} (\mathfrak{L}_{\underline{\theta}, N}(f) -
\mathfrak{L}^{R}_{2 \omega, N}(f)) = 0.
\end{equation}
\end{theorem}
This says that the values of $\mathfrak{L}_{\underline\theta, N}$
are deterministic in the large $N$ limit. However,
they do not converge to a single value as the one
for the irrational rotation does not.
Using either Theorem~\ref{thm:main4} or easier
Theorem~\ref{thm:evskew}, one can show that the
gap distribution of the eigenvalues indeed
obeys the distribution shown in Figure~\ref{fig:1} and
\ref{fig:2}. The Diophantine
assumption \eqref{eq:conddiop} is necessary,
I sketch an argument in Remark~\ref{rem:liouville}.
Furthermore, it should be noted that Lebesgue
almost every $\omega$ satisfies \eqref{eq:conddiop}.
In this sense the case $k = 2$ is of intermediate
disorder, one has pure point spectrum with exponentially
decaying eigenfunctions, but one does not have sufficient
independence to obtain Poisson statistics.
The definition of the Laplace functional given here
is different from the one usually given in the theory
of point processes. There, one does not introduce
averaging over the unit circle by hand, but this comes
from the points $x_n$ being defined on some probability
space. In Section~\ref{sec:skew}, we will see that
our Verblunsky coefficients are defined on a probability
space, and that averaging over it in particular contains
the $\theta$ average. Hence, the name Laplace functional
is justified.
\begin{remark}\label{rem:liouville}
Assume that for coprime integers $p,q$, $N$
very large, and $\delta > 0$ a small parameter, we have
that $|\omega-\frac{p}{q}|\leq \frac{1}{N^{3 + \delta}}$.
Then for $1 \leq n \leq N$, we have that
\begin{equation}
\left|\alpha_n - \lambda \mathrm{e}^{2\pi\mathrm{i} \frac{p n^2}{q}}
\right|\leq\frac{1}{N^{1 + \frac{\delta}{2}}}.
\end{equation}
Since the Verblunsky coefficients $\lambda \mathrm{e}^{2\pi\mathrm{i} \frac{p n^2}{q}}$
are $q$-periodic, the corresponding zeros of the
paraorthogonal polynomials are clock-spaced, so
of size $\frac{1}{N}$, whereas the points
$\{2 n\omega\pmod{1}\}_{n=1}^{N}$
are all in a $\frac{1}{N^{2 + \delta}}$
neighborhood of the points $\{\frac{\ell}{q}\}_{\ell=1}^{q}$.
These two behaviors are clearly incompatible, and
thus Theorem~\ref{thm:main4} cannot hold for
Liouville frequencies.
\end{remark}
Let me now outline the rest of the content of
the paper. Section~\ref{sec:pfmain1} discusses the
basic theory of half-line CMV matrices and gives
the proof of Theorem~\ref{thm:main1}. Then Section~\ref{sec:excmv}
introduces extended CMV operators, so ones
defined on the whole-line, discusses restrictions
of these, defines the Green's function,
and derives useful formulas relating determinants
of CMV matrices to transfer matrices. This discussion
is somewhat more complicated than the case of
Schr\"odinger operators. Section~\ref{sec:ueCMV}
combines the formulas from the previous section
with the ones for ergodic CMV matrices.
In Section~\ref{sec:skew}, CMV matrices with built-in
rotational invariance are discussed and
Theorem~\ref{thm:main2} is proven.
In Section~\ref{sec:evstat}, we prove Theorem~\ref{thm:main4}
relying on results from Sections~\ref{sec:existtest}
and \ref{sec:greendecays}. Basically, Section~\ref{sec:greendecays}
improves the bounds on decay of the Green's function
obtained in Section~\ref{sec:ueCMV} from
unique ergodicity by using quantitative recurrence
results for the skew-shift discussed in Appendix~\ref{sec:dynskew}.
Section~\ref{sec:existtest} shows how to exploit
Section~\ref{sec:greendecays} to obtain good
test functions.
\section{A first look at the CMV matrix}
\label{sec:pfmain1}
In this section, we take a look at half-line CMV matrices
and provide a proof of Theorem~\ref{thm:main1}. In the following
sections, we will discuss whole line CMV matrices in more
details. Although most results in this section will be reproven
in later parts, I have included it, since it is closed
to the notation of \cite{opuc1,opuc2}.
Let $\{\alpha_n\}_{n=0}^{\infty}$ be a sequence of Verblunsky
coefficients. Define $\rho_n = (1 - |\alpha_n|^2)^{\frac{1}{2}}$
and the unitary matrices
\begin{equation}
\Theta_n = \begin{pmatrix} \overline{\alpha_n} & \rho_n \\ \rho_n
& - \alpha_n \end{pmatrix}.
\end{equation}
Define the operators $\mathcal{L}_+, \mathcal{M}_+$ by
\begin{equation}
\mathcal{L}_+ = \begin{pmatrix} \Theta_0 \\ & \Theta_2 \\ & & \ddots
\end{pmatrix},\quad
\mathcal{M}_+ = \begin{pmatrix} 1 \\ & \Theta_1 \\ & & \ddots
\end{pmatrix}
\end{equation}
where $1$ represents the identity $1 \times 1$ matrix.
The CMV matrix is then defined by $\mathcal{C} = \mathcal{L}_+
\mathcal{M}_+$ which will be five-diagonal and unitary. Its importance
comes from that the measure $\mu$ associated to the
Verblunsky coefficients $\{\alpha_n\}_{n=0}^{\infty}$ is
the spectral measure of $\delta_0$ with respect to
$\mathcal{C}$, so one has
\begin{equation}
\int_{\partial{\mathbb D}} z^n d\mu(z) =
\spr{\delta_0}{\mathcal{C}^n \delta_0}.
\end{equation}
We denote by $\mathrm{supp}_{\mathrm{ess}}(\mu)$ the
essential support of the measure $\mu$, that is the
support of $\mu$ with point masses removed.
\begin{lemma}\label{lem:translatealpha}
Define $\tilde\alpha_n = \alpha_{n+1}$. Let $\tilde{\mu}$ be
the measure corresponding to $\{\tilde\alpha_n\}_{n=0}^{\infty}$.
Then
\begin{equation}
\mathrm{supp}_{\mathrm{ess}}(\tilde{\mu})
= \mathrm{supp}_{\mathrm{ess}}(\mu).
\end{equation}
\end{lemma}
\begin{proof}
Clearly $\mathrm{supp}_{\mathrm{ess}}(\mu) =
\sigma_{\mathrm{ess}}(\mathcal{C})$. Let $S$ be the backward
shift on $\ell^2({\mathbb N})$. Then $\mathcal{C}$ and
$S^* \mathcal{\tilde{C}} S$ differ by a finite rank operator.
The claim follows.
\end{proof}
A similar proof implies that for all the translates
$\alpha^{\ell}_n = \alpha_{n+\ell}$ the corresponding
CMV matrices have the same essential spectrum.
Hence, for Verblunsky coefficients given by
\eqref{eq:defVeromegank}, one obtains that the family
of Verblunsky coefficients given by
\begin{equation}
\alpha^{\ell}_n = \lambda \exp\left(2\pi\mathrm{i}
\left(\omega n^k + \sum_{j=0}^{k-1}
\binom{k}{j} \omega \ell^{k-j} \cdot n^j\right)\right)
\end{equation}
have the same essential spectrum. Define
for $y \in [0,1]^k$ a family of Verblunsky
coefficients by
\begin{equation}
\tilde{\alpha}_{y, n} = \lambda \exp\left(2\pi\mathrm{i}
\left(\omega n^k + \sum_{j=0}^{k-1}
y_j \cdot n^j\right)\right).
\end{equation}
\begin{lemma}\label{lem:sigmaCy}
We have for any $y\in [0,1]^k$ that
\begin{equation}
\sigma_{\mathrm{ess}}(\mathcal{C}) =
\sigma_{\mathrm{ess}}(\widetilde{\mathcal{C}}_{y})
\end{equation}
\end{lemma}
\begin{proof}
Given $y$, there exists a sequence $\ell_s$ such that
\[
\binom{k}{j} \omega \ell_s^{k-j} \to y_j
\]
for $0\leq j\leq k-1$ as $s\to\infty$
(see Theorem~2.2 and Lemma~2.3 in \cite{krho1}).
By strong convergence, one thus obtains that
\[
\sigma_{\mathrm{ess}}(\mathcal{C}) \supseteq
\sigma_{\mathrm{ess}}(\widetilde{\mathcal{C}}_{y}).
\]
The other inclusion can be proven in a similar way.
\end{proof}
Results similar to Lemma~\ref{lem:sigmaCy} have been
discussed in \cite{lsess}.
For the proof of Theorem~\ref{thm:main1},
we will also need
\begin{proposition}\label{prop:rotatealpha}
Define Verblunsky coefficients by
$\tilde{\alpha}_n = \mathrm{e}^{2\pi\mathrm{i} \eta n} \alpha_n$.
Then
\begin{equation}
\mathrm{supp}(\tilde\mu) = \mathrm{e}^{-2\pi\mathrm{i} \eta}
\mathrm{supp}(\mu).
\end{equation}
\end{proposition}
\begin{proof}
This follows from the formulas in
Appendix~A.H. in \cite{opuc2}. I will
also give another proof in Section~\ref{sec:skew}.
\end{proof}
Given $y \in [0,1]^k$ and $\eta \in [0,1]$ define
\begin{equation}
\hat{y}_j = \begin{cases} y_j,&j=0, 2 \leq j \leq k-1;\\
y_1 + \eta,& j=1.\end{cases}
\end{equation}
Proposition~\ref{prop:rotatealpha} shows that
\begin{equation}
\sigma_{\mathrm{ess}}(\widetilde{\mathcal{C}}_{y}) =
\mathrm{e}^{- 2\pi\mathrm{i} \eta}
\sigma_{\mathrm{ess}}(\widetilde{\mathcal{C}}_{\hat{y}}).
\end{equation}
Having this, we are now ready for
\begin{proof}[Proof of Theorem~\ref{thm:main1}]
The results discussed so far imply that
$\sigma_{\mathrm{ess}}(\mathcal{C})$ is a non-empty,
rotationally invariant, subset of $\partial{\mathbb D}$.
Hence, we must have
\[
\sigma_{\mathrm{ess}}(\mathcal{C}) = \partial{\mathbb D}
\]
Since also $\sigma_{\mathrm{ess}}(\mathcal{C})\subseteq
\sigma(\mathcal{C})\subseteq\partial{\mathbb D}$,
the claim follows.
\end{proof}
\section{Extended CMV operators}
\label{sec:excmv}
In this section, we introduce extended CMV operators
and discuss their properties that will be useful to
us. See also \cite{gzweyl} and Section~10.5 in \cite{opuc2}
for discussions from different viewpoints.
Let now $\{\alpha_n\}_{n\in{\mathbb Z}}$ be a bi-infinite
sequence of Verblunsky coefficients, i.e.
$\alpha_n \in {\mathbb D}$ although we will discuss setting
certain $\alpha_{n}$ to values in $\overline{\mathbb D}$ below.
Recall that $\rho_n = (1-|\alpha_n|^2)^{\frac{1}{2}}$ and
\begin{equation}
\Theta_n = \begin{pmatrix} \overline{\alpha_n} & \rho_n \\
\rho_n & - \alpha_n \end{pmatrix}
\end{equation}
viewed as acting on $\ell^2(\{n,n+1\})$. Define
\begin{equation}
\mathcal{L} = \bigoplus_{n\text{ even}} \Theta_n,\quad
\mathcal{M} = \bigoplus_{n\text{ odd}} \Theta_n
\end{equation}
and the extended CMV operator $\mathcal{E} = \mathcal{L}
\cdot \mathcal{M}$. We note
\begin{lemma}
$\mathcal{E}$, $\mathcal{L}$, and $\mathcal{M}$ are
unitary operators $\ell^2({\mathbb Z})\to\ell^2({\mathbb Z})$. Furthermore,
$\mathcal{L}$ leaves the subspaces $\ell^2(\{n,n+1\})$
for $n$ even invariant, whereas $\mathcal{L}$ does this for
$n$ odd.
\end{lemma}
We will now discuss various restrictions of CMV
operators. First denote by $P^{[a,b]}$ the projection
$\ell^2({\mathbb Z}) \to \ell^2([a,b])$. We define
\begin{equation}
X^{[a,b]} = (P^{[a,b]})^* X P^{[a,b]}
\end{equation}
for $X\in\{\mathcal{E},\mathcal{M},\mathcal{L}\}$.
\begin{lemma}
$\mathcal{E}^{[a,b]} = \mathcal{L}^{[a,b]} \mathcal{M}^{[a,b]}$.
\end{lemma}
\begin{proof}
Compute.
\end{proof}
It is easy to check that the operator $\mathcal{E}^{[a,b]}$
will no longer be unitary, but it will still be an useful
object. Let now $\beta \in \partial{\mathbb D}$ and $a\in{\mathbb Z}$
and consider the modified Verblunsky coefficients
\begin{equation}
\tilde{\alpha}_n = \begin{cases} \alpha_n,& n\neq a\\
\beta,& n=a.\end{cases}
\end{equation}
We then have that $\widetilde{\mathcal{E}}$,
$\widetilde{\mathcal{L}}$, and
$\widetilde{\mathcal{M}}$ leave the
spaces $\ell^2(\{a+1, a+2, \dots\})$ and $\ell^2(\{\dots, a-1,a\})$
invariant. In particular, we can define unitary
restrictions
\begin{equation}
\mathcal{E}^{[a+1, \infty)}_{\beta, \bullet} = P^{[a+1,\infty)}
\widetilde{\mathcal E} P^{[a+1, \infty)},\quad
\mathcal{E}^{(-\infty,a]}_{\bullet, \beta} = P^{(-\infty,a]}
\widetilde{\mathcal E} P^{(-\infty,a]}.
\end{equation}
\begin{lemma}\label{lem:relationC}
Let $\mathcal{C}$ be the CMV operator with Verblunksy
coefficients $\{\alpha_n\}_{n=0}^{\infty}$. Then
\begin{equation}
\mathcal{C} = \mathcal{E}^{[0,\infty)}_{1, \bullet}.
\end{equation}
Denote by $R$ the identification $\ell^2(\{\dots, -2,-1\})$
with $\ell^2(\{0,1,2,\dots\})$ and
by $\mathcal{C}_-$ the CMV operator with Verblunsky
coefficients $\{-\overline{\alpha}_{-n-1}\}_{n=0}^{\infty}$.
Then
\begin{equation}
\mathcal{C}_- = R \mathcal{E}^{(-\infty,-1]}_{\bullet, 1} R^{*}.
\end{equation}
\end{lemma}
\begin{proof}
These are computations.
\end{proof}
We will now consider restrictions to intervals.
So let $a < b$ be integers, and $\beta,\gamma \in
\partial{\mathbb D}$. Define a sequence of Verblunsky
coefficients
\begin{equation}
\tilde{\alpha}_n = \begin{cases} \beta,&n=a;\\
\gamma,&n=b;\\
\alpha_n,& n\notin\{a,b\} \end{cases}.
\end{equation}
We then define the operator
\begin{equation}
\mathcal{E}^{[a+1,b]}_{\beta, \gamma}
= P^{[a+1,b]} \widetilde{\mathcal{E}} P^{[a+1,b]}.
\end{equation}
Of course, this definition makes sense
for $\beta,\gamma \in \overline{{\mathbb D}}$ and $a = - \infty$
or $b = \infty$. Furthermore, we write
$\bullet$ if we leave $\alpha_a$ or $\alpha_b$
unchanged to match the previous definition.
$\beta,\gamma\in\partial{\mathbb D}$ should be thought of
as boundary conditions.
\begin{lemma}
If $\beta,\gamma\in\partial{\mathbb D}$ then
$\mathcal{E}^{[a,b]}_{\beta,\gamma}$,
$\mathcal{L}^{[a,b]}_{\beta,\gamma}$, and
$\mathcal{M}^{[a,b]}_{\beta,\gamma}$ are unitary.
\end{lemma}
Since the equation $\mathcal{E} \psi = z \psi$
is equivalent to $(z \mathcal{L}^{\ast} - \mathcal{M}) \psi = 0$.
We note for further reference
\begin{lemma}\label{lem:tridiag}
The matrix $A = z (\mathcal{L}^{[a,b]}_{\beta,\gamma})^{*}
- \mathcal{M}^{[a,b]}_{\beta,\gamma}$ is tridiagonal.
Write $A = \{A_{i,j}\}_{a\leq i,j \leq b}$.
Then we have that
\begin{equation}
A_{j,j} = \begin{cases}
z \alpha_j + \alpha_{j-1},&j\text{ even}\\
- z \overline{\alpha_{j-1}}
- \overline{\alpha_{j}},&j\text{ odd},\end{cases}\quad
A_{j+1, j} = A_{j,j+1} = \tilde{\rho}_j =
\begin{cases} z \rho_j,&j\text{ even},\\
- \rho_{j},&j\text{ odd}.\end{cases}
\end{equation}
\end{lemma}
Let $z\in{\mathbb C}$, $\beta,\gamma\in\partial{D}$,
$a\leq k,\ell\leq b$, then the Green's function is
defined by
\begin{equation}
G^{[a,b]}_{\beta,\gamma}(z; k,\ell)=
\spr{\delta_k}{(z \left(\mathcal{L}^{[a,b]}_{\beta,\gamma})^{*}
- \mathcal{M}^{[a,b]}_{\beta,\gamma}\right)^{-1} \delta_\ell}.
\end{equation}
Our goal now will be to provide a formula
for the Green's function in terms of quantities
that are easier to analyze, like the formula
for the Green's function of Schr\"odinger operators
in term of orthogonal polynomials, respectively
entries of the transfer matrix.
We define
\begin{align}
\Phi^{[a,b]}_{\beta,\gamma}(z)
&= \det\left(z - \mathcal{E}^{[a,b]}_{\beta,\gamma}\right) \\
\nonumber &= \det\left(
z (\mathcal{L}^{[a,b]}_{\beta,\gamma} )^*
- \mathcal{M}^{[a,b]}_{\beta,\gamma}
\right) \cdot \det((\mathcal{L}^{[a,b]}_{\beta,\gamma} )^*)
\end{align}
and
\begin{equation}
\varphi^{[a,b]}_{\beta,\gamma}(z) = (\rho_a \cdots \rho_{b})^{-1}
\Phi^{[a,b]}_{\beta,\gamma}(z).
\end{equation}
\begin{lemma}
Let $\Phi_n(z)$ be defined as in \eqref{eq:defPhin}.
Then
\begin{equation}
\Phi_{n}(z) = \Phi^{[0,n-1]}_{1,\bullet}(z).
\end{equation}
\end{lemma}
\begin{proof}
Proposition~3.4. in \cite{scmv5} states
\[
\Phi_n(z) = \det(z - \mathcal{E}^{[0,n-1]}_{1, \bullet}).
\]
The claim follows.
\end{proof}
We also introduce the Aleksandrov polynomials $\Phi^{\beta}_n(z)$
by applying the recursion \eqref{eq:defPhin} to
the Verblunsky coefficients $\{\beta \alpha_n\}_{n=0}^{\infty}$.
In particular, the polynomial of the second
kind is defined by
\begin{equation}
\Psi_n(z) = \Phi^{-1}_n(z).
\end{equation}
We have that (Theorem~9.5. in \cite{s1foot})
\begin{lemma}
We have
\begin{equation}
\Phi^{\beta}_n(z) = \Phi_{\beta,\bullet}^{[0,n-1]}(z)
\end{equation}
and
\begin{equation}
\Phi^{\beta}_n(z;\gamma) =
\Phi_{\beta,\gamma}^{[0,n-1]}(z)
\end{equation}
\end{lemma}
With these formulas, we obtain the following equality
for the absolute value of the Green's function. It would
be possible to derive an equality for the Green's function
but one would need distinguish between 4 cases depending
on if $a$ or $b$ is even or odd.
\begin{proposition}\label{prop:greenaspolynomial}
Let $z\in{\mathbb C}$, $\beta,\gamma\in\partial{\mathbb D}$,
and $a \leq k \leq \ell \leq b$. Then
\begin{equation}
\left| G^{[a, b]}_{\beta,\gamma} (z; k, \ell)\right|
= \frac{1}{\rho_k \rho_{\ell}}
\left| \frac{\varphi_{\beta, \bullet}^{[a,k-1]}(z)
\varphi_{\bullet, \gamma}^{[\ell+1, b]}(z)}
{\varphi_{\beta,\gamma}^{[a,b]}(z)}\right|
\end{equation}
\end{proposition}
\begin{proof}
By Cramer's rule and Lemma~\ref{lem:tridiag},
we thus obtain
\[
\left|G^{[a, b]}_{\beta,\gamma} (z; k, \ell)\right|
= \tilde{\rho}_{k+1} \cdots \tilde{\rho}_{\ell - 1}
\left|\frac{\Phi_{\beta, \bullet}^{[a,k-1]}(z)
\Phi_{\bullet, \gamma}^{[\ell+1, b]}(z)}
{\Phi_{\beta,\gamma}^{[a,b]}(z)}\right|
\]
The claim now follows from the definition of $\varphi$.
\end{proof}
This formula is more awkward than the one for
Schr\"odinger operators, since it involves
three different type of polynomials whereas the
one for Schr\"odinger operators only has one
(see (2.7) in \cite{bbook}). Nevertheless it
is useful in exactly the same way.
We now give the relation of the Green's function
to solution of our equation.
\begin{lemma}\label{lem:green2sol}
Let $\psi$ solve $\mathcal{E} \psi = z \psi$. Then
for $a < n < b$
\begin{align}
\psi(n) & = G^{[a,b]}_{\beta,\gamma}(z;n,a)
\begin{cases} (z \overline{\beta} - \alpha_a) \psi(a)
-\rho_a \psi(a+1),& a\text{ even};\\
(z \alpha_a - \beta) \psi(a) + z \rho_a \psi(a+1),&a\text{ odd}
\end{cases}\\
\nonumber &+ G^{[a,b]}_{\beta,\gamma}(z;n, b)
\begin{cases} (z \overline{\gamma} - \alpha_b) \psi(b)
-\rho_b \psi(b-1),& b\text{ even};\\
(z \alpha_b - \gamma) \psi(b)
+ z \rho_{b-1} \psi(b-1),&b\text{ odd}
\end{cases}
\end{align}
\end{lemma}
\begin{proof}
With $A = (z(\mathcal{L}^{[a,b]}_{\beta,\gamma})^{\ast} -
\mathcal{M}^{[a,b]}_{\beta,\gamma})$, we have
\[
\varphi(n) = \spr{A^{-1} \delta_n, A \varphi}.
\]
Since, $(z(\mathcal{L})^{\ast} - \mathcal{M} )\varphi=0$,
we have that for $a+1 \leq n \leq b$ also
\[
A \varphi (n) = 0.
\]
The claim now follows by evaluating this
expression for $n \in \{a,b\}$.
\end{proof}
Our next goal will be to introduce transfer
matrices and related them to the determinants
defined above. We begin with the one-step
transfer matrix
\begin{equation}
A_z(\alpha) = \frac{1}{(1-|\alpha|^2)^{\frac{1}{2}}}
\begin{pmatrix} z & - \overline{\alpha} \\
-\alpha z & 1 \end{pmatrix}.
\end{equation}
We define the transfer matrix by
\begin{equation}
T^{[a,b]}(z) = A_z(\alpha_{b}) \cdots A_z(\alpha_a).
\end{equation}
\begin{lemma}
We have that
\begin{equation}
T^{[a,b]}(z) = \frac{1}{2} \begin{pmatrix}
\varphi^{[a,b]}_{1,\bullet}(z) +
\varphi^{[a,b]}_{-1,\bullet}(z) &
\varphi^{[a,b]}_{1,\bullet}(z) -
\varphi^{[a,b]}_{-1,\bullet}(z) \\
(\varphi^{[a,b]}_{1,\bullet})^*(z) -
(\varphi^{[a,b]}_{-1,\bullet})^*(z) &
(\varphi^{[a,b]}_{1,\bullet})^*(z)
+ (\varphi^{[a,b]}_{-1,\bullet})^*(z)\end{pmatrix}.
\end{equation}
where $(\varphi^{[a,b]}_{\beta,\gamma})^*(z)
= z^{b-a+1} \overline{\varphi^{[a,b]}_{\beta,\gamma}(\overline{z}^{-1})}$.
\end{lemma}
\begin{proof}
The $T_n(z)$ in \cite{s1foot} is $T^{[0,n-1]}(z)$
in our notation. We have that
\[
T_n(z) = \frac{1}{2} \begin{pmatrix}
\varphi_n(z) + \psi_n(z) & \varphi_n(z) - \psi_n(z) \\
\varphi_n^*(z) - \psi_n^*(z) & \varphi_n^*(z)
+ \psi_n^*(z)\end{pmatrix}.
\]
It follows that
\[
T^{[0,n-1]}(z) = \frac{1}{2} \begin{pmatrix}
\varphi^{[0,n-1]}_{1,\bullet}(z) +
\varphi^{[0,n-1]}_{-1,\bullet}(z) &
\varphi^{[0,n-1]}_{1,\bullet}(z) -
\varphi^{[0,n-1]}_{-1,\bullet}(z) \\
(\varphi^{[0,n-1]}_{1,\bullet})^*(z) -
(\varphi^{[0,n-1]}_{-1,\bullet})^*(z) &
(\varphi^{[0,n-1]}_{1,\bullet})^*(z)+
(\varphi^{[0,n-1]}_{-1,\bullet})^*(z)\end{pmatrix}.
\]
The claim follows using translation invariance.
\end{proof}
We thus obtain that
\begin{corollary}
We have that
\begin{equation}\label{eq:varphibetabullet}
\begin{pmatrix} \varphi^{[a,b]}_{\beta,\bullet}(z)\\
(\varphi^{[a,b]}_{\beta,\bullet})^{*}(z) \end{pmatrix} =
T^{[a,b]}(z) \begin{pmatrix} 1 \\ \overline{\beta} \end{pmatrix}
\end{equation}
and
\begin{equation}\label{eq:varphibetagamma}
\varphi^{[a,b]}_{\beta,\gamma}(z) =
\frac{1}{\rho_b}
\spr{\begin{pmatrix} z \\ - \overline{\gamma} \end{pmatrix}}
{T^{[a,b-1]}(z) \begin{pmatrix} 1 \\ \overline{\beta} \end{pmatrix}}.
\end{equation}
\end{corollary}
\begin{proof}
The first equation is (3.2.26) in \cite{opuc1}.
For the second equation, we have that
\[
\Phi^{[a,b]}_{\beta,\gamma}(z) = z
\Phi^{[a,b-1]}_{\beta,\bullet}(z) - \overline{\gamma}
(\Phi^{[a,b-1]}_{\beta,\bullet})^*(z).
\]
We thus have that
\[
\varphi^{[a,b]}_{\beta,\gamma}(z) =
\frac{1}{\rho_b} \left( z
\varphi^{[a,b-1]}_{\beta,\bullet}(z) - \overline{\gamma}
(\varphi^{[a,b-1]}_{\beta,\bullet})^*(z)
\right),
\]
which implies the second equation by the first one.
\end{proof}
There is one final object, we need to identify
$\varphi^{[a,b]}_{\bullet, \gamma}(z)$. We employ
the same strategy as we used in Lemma~\ref{lem:relationC}
to identify $\mathcal{E}^{(-\infty,0]}_{\bullet,\gamma}$.
Let
\begin{equation}
\tilde{\alpha}_{n} = \begin{cases} \gamma, & n = -1;\\
- \overline{\alpha}_{b-n}, &n \geq 0.\end{cases}
\end{equation}
Then we have that
\begin{equation}
\varphi^{[a,b]}_{\bullet, \gamma}(z) =
\tilde{\varphi}^{[0,b-a-1]}_{\gamma,\bullet}(z).
\end{equation}
\begin{lemma}
We have htat
\begin{equation}\label{eq:varphibulletgamma}
\begin{pmatrix} \varphi^{[a,b]}_{\bullet,\gamma}(z) \\
(\varphi^{[a,b]}_{\bullet,\gamma})^*(z) \end{pmatrix} =
\begin{pmatrix} - \frac{1}{z} & 0 \\ 0 & 1 \end{pmatrix}
(T^{[a,b]}(z))^t \begin{pmatrix} -z\\\overline{\gamma} \end{pmatrix}.
\end{equation}
\end{lemma}
\begin{proof}
We have that
\[
\begin{pmatrix} - \frac{1}{z} & 0 \\ 0 & 1 \end{pmatrix}
A_z(-\overline{\alpha})^{t} \begin{pmatrix} -z & 0 \\ 0 & 1 \end{pmatrix} = A_z(\alpha).
\]
From this the claim follows.
\end{proof}
\section{Strictly ergodic CMV matrices}
\label{sec:ueCMV}
In this section, we will consider families of CMV operators.
This has the advantage that certain formulas will simplify,
when viewed probabilistically. Also strict ergodicity
simplifies certain statements not available in the ergodic
case, in particular \cite{furman}.
Let $\Omega$ be a compact metric space,
$T:\Omega\to\Omega$ a uniquely ergodic
and minimal homeomorphism,
and $\mu$ the unique $T$-invariant probability measure.
We call $(\Omega,\mu,T)$ strictly ergodic in this case.
For a continuous function $f:\Omega\to{\mathbb D}$, we define
the family of Verblunsky coefficients
\begin{equation}
\alpha_{\omega, n} = f(T^n \omega).
\end{equation}
We denote by $\mathcal{E}_{\omega}, \dots$ the associated
objects.
The main example to keep in mind is the $k$-dimensional
skew-shift with $\Omega = {\mathbb T}^k = ({\mathbb R}/{\mathbb Z})^k$
\begin{equation}\label{eq:defkskew}
(T x)_{\ell} = \begin{cases} x_1 + \omega, & \ell = 1; \\
x_{\ell} + x_{\ell - 1}, &2 \leq \ell \leq k.\end{cases}
\end{equation}
One can then show by induction that
\begin{equation}
(T^n x)_{\ell} = \binom{n}{\ell} \omega + \binom{n}{\ell -1} x_1
+ \dots + \binom{n}{0} x_{\ell}.
\end{equation}
This map is strictly ergodic, see Proposition~4.7.4. in \cite{brinstuck}.
Then one can realize the Verblunsky coefficient
from the introduction as $\alpha_{x, n}$
for $f(x) = \lambda \mathrm{e}^{2\pi\mathrm{i} x_k}$ and a particular
choice of $x$.
We now return to our study of the general case
of uniquely ergodic and minimal CMV matrices.
\begin{lemma}
We have that $\mathcal{E}_{T x} = (S^{\ast} \mathcal{E}_x S)^{t}$,
where $S$ is the usual forward shift on $\ell^2({\mathbb Z})$. In
particular for any $x,y\in\Omega$
\begin{equation}
\sigma(\mathcal{E}_x) = \sigma(\mathcal{E}_y).
\end{equation}
\end{lemma}
\begin{proof}
The first claim is algebraic. The second
claim follows as Lemma~\ref{lem:sigmaCy}.
\end{proof}
For $n\geq 1$, we define the $n$-step (forward) transfer matrix by
\begin{equation}\label{eq:defTxn}
T_{x;n}(z) = A(\alpha_{x,n-1}, z) \cdots A(\alpha_{x,0}, z).
\end{equation}
We note that $T_{x;n}(z) = T^{[0,n-1]}_{x}(z)$ in the notation
of the previous section, and that also
$T^{[a,b]}_x(z) = T_{T^{a} x; b-a+1}(z)$.
The Lyapunov exponent is defined by
\begin{equation}
L(z) = \lim_{n\to\infty} \frac{1}{n} \int_{{\mathbb T}^k}
\log\|T_{n,x}(z)\| dx.
\end{equation}
We collect its properties
\begin{proposition}\label{prop:proplyap}
Let $(\Omega,\mu,T)$ be strictly ergodic and
$z\in\partial{\mathbb D}$.
\begin{enumerate}
\item $L(z) \geq 0$.
\item For almost-every $x\in{\mathbb T}^k$, we have as $n\to\infty$
that
\begin{equation}
\frac{1}{n} \log\|T_{x; n}(z)\| \to L(z).
\end{equation}
\item For every $\varepsilon > 0$, there exists $N$ such that
for $n \geq N$ and $x\in{\mathbb T}^K$ we have
\begin{equation}
\frac{1}{n} \log\|T_{x; n}(z)\| \leq L(z) + \varepsilon.
\end{equation}
\end{enumerate}
\end{proposition}
\begin{proof}
(i) follows from $\det(A(\alpha,z)) = z$.
(ii) is the subadditive ergodic theorem (see Corollary 10.5.25
in \cite{opuc2}).
(iii) is Furman's strengthening for uniquely
ergodic transformations \cite{furman}.
\end{proof}
The right extension of \eqref{eq:defTxn}
for negative numbers is
\begin{equation}
T_{x; - n}(z) = \begin{pmatrix} -\frac{1}{z} & 0 \\ 0 & 1
\end{pmatrix} A(-\overline{\alpha_{x, -1}}, z) \cdots
A(-\overline{\alpha_{x,-n}}, z) \begin{pmatrix}
-z & 0 \\ 0 & 1 \end{pmatrix}
\end{equation}
(where $n\geq 0$). This can be seen from
\eqref{eq:varphibulletgamma}. In particular,
one has
\begin{equation}
L(z) = \lim_{n\to\infty} \frac{1}{n} \int_{{\mathbb T}^k}
\log\|T_{x; -n}(z)\|dx.
\end{equation}
\begin{lemma}
Let $(\Omega,\mu,T)$ be strictly ergodic
and $\varepsilon > 0$.
There exists $C > 1$ such that for $n \geq 1$
and $\beta,\gamma \in \partial{\mathbb D}$,
we have for $0\leq k \leq \ell \leq n-1$ that
\begin{equation}
|G_{x; \beta,\gamma}^{[0,n-1]}(z; k,\ell)| \leq
C \frac{\mathrm{e}^{ (L(z) + \varepsilon) (k + n - 1 - \ell)}}
{|\varphi^{[0,n-1]}_{x; \beta,\gamma}(z)|}.
\end{equation}
\end{lemma}
\begin{proof}
By (iii) of Proposition~\ref{prop:proplyap}, there exists $c \geq 1$
such that for any $x\in\Omega$ and $n\geq 1$, we have
\[
\|T_{x; n}(z)\| \leq c \mathrm{e}^{(L(z) + \varepsilon) n}.
\]
By \eqref{eq:varphibetabullet} and \eqref{eq:varphibulletgamma},
we obtain that the numerator in Proposition~\ref{prop:greenaspolynomial}
is bounded by
\[
c_2 \cdot \mathrm{e}^{(L(z) + \varepsilon) (n-1 - \ell + k)}.
\]
The claim follows.
\end{proof}
In particular, we obtain the important theorem
\begin{theorem}\label{thm:uegreen}
Let $(\Omega,\mu,T)$ be strictly ergodic,
$m \in (0, L(E))$, $\delta > 0$,
and $\beta_0, \gamma_0\in\partial{\mathbb D}$.
Then for $n$ large enough, there exists $\Omega_n$
satisfying $\mu(\Omega_n) \geq 1 - \delta$ and
for $x \in \Omega_n$ there exists
\begin{equation}
\beta \in \{-\beta_0,\beta_0\},\quad
\gamma\in\{-\gamma_0,\gamma_0\}
\end{equation}
such that
for $\frac{n}{3} \leq \ell \leq \frac{2n}{3}$
and $k \in \{0,n-1\}$
\begin{equation}
|G_{x; \beta,\gamma}^{[0,n-1]}(z; k,\ell)| \leq
\mathrm{e}^{- m |k - \ell|}.
\end{equation}
\end{theorem}
\begin{proof}
By \eqref{eq:varphibetagamma}, we have that
\[
\begin{pmatrix} \varphi^{[0,n-1]}_{x;\beta,\gamma}(z) &
\varphi^{[0,n-1]}_{x;-\beta,\gamma}(z) \\
\varphi^{[0,n-1]}_{x;\beta,-\gamma}(z) &
\varphi^{[0,n-1]}_{x;-\beta,-\gamma}(z) \end{pmatrix}
= \begin{pmatrix} z & -\overline{\gamma} \\ z & \overline{\gamma}
\end{pmatrix}
T_{x,n}(z) \begin{pmatrix} 1 & 1 \\ \beta & -\beta \end{pmatrix}
\]
Since for almost every $x$
$\frac{1}{n}\log\|T_{x,n}(z)\| \geq L(z) (1 - \varepsilon)$
for $n$ large enough, the claim
follows.
\end{proof}
\section{Rotationally invariance and the proof of Theorem~\ref{thm:main2}}
\label{sec:skew}
We begin this section by investigating what
happens if one rotates the Verblunsky coefficients,
which is essentially what we used to prove Theorem~\ref{thm:main1}.
We have the following important proposition
\begin{proposition}\label{prop:rotatealphas}
Let $\beta,\gamma\in\overline{\mathbb D}$, $a<b$ integers, and
$x,y\in{\mathbb T}$ and define
\begin{equation}
\tilde{\alpha}_n = e(n x+y) \alpha_n,\quad
\tilde{\beta} = e((a-1) x+ y) \beta,\quad
\tilde{\gamma} = e(b x + y) \gamma.
\end{equation}
Then $\mathcal{E}^{[a,b]}_{\beta,\gamma}$
and $e(x) \widetilde{\mathcal{E}}^{[a,b]}_{\tilde{\beta},\tilde{\gamma}}$
are unitarily equivalent.
\end{proposition}
Here and in the following, we abbreviate
$e(x) = \mathrm{e}^{2\pi\mathrm{i} x}$.
We will prove this proposition in the case of $a$ and
$b$ finite. It is interesting if it holds for $a,b$
possibly infinite. An inspection of the proof of
Proposition~\ref{prop:rotatealphas} shows that it
also holds for whole line CMV operators with pure
point spectrum. In particular, it implies that
in the case $k = 2$, all the operators $\mathcal{E}_x$
defined by the skew-shift are unitarily equivalent.
Since the Jitomirskaya--Simon \cite{js} argument
applies in our case, all the $\mathcal{E}_x$ have purely
singular continuous spectrum.
For the proof of this proposition, we need the following
lemma
\begin{lemma}\label{lem:rotatealpha}
Pick some $u_a\in\partial{\mathbb D}$ and define a sequence
recursively by
\begin{equation}
u_{n} = \begin{cases} u_{n-1} e(-(n-1) x - y),&n\text{ even};\\
u_{n-1} e((n-1) x+y),&n\text{ odd}.\end{cases}
\end{equation}
Furthermore, we define the multiplication operators
\begin{equation}
U \psi(n) = u_n \psi(n),\quad V \psi(n) = \begin{cases}
u_{n-1} \psi(n), & n\text{ even};\\
u_{n-1} e(-x) \psi(n), & n\text{ odd}\end{cases}.
\end{equation}
Then for $\tilde{z} = e(-x) z$
\begin{equation}
\left(\tilde{z} (\widetilde{\mathcal{L}}^{[a,b]}_{\tilde\beta,\tilde\gamma})^*
- \widetilde{\mathcal{M}}^{[a,b]}_{\tilde\beta,\tilde\gamma}\right) U =
V \left(z (\mathcal{L}^{[a,b]}_{\beta,\gamma})^*
- \mathcal{M}^{[a,b]}_{\beta,\gamma}\right).
\end{equation}
\end{lemma}
\begin{proof}
A computation shows for $n$ even that
\[
\tilde{z} \tilde{\alpha}_n + \tilde{\alpha}_{n-1} = e((n-1)x+y) (z\alpha_n+\alpha_{n-1})
\]
and for $n$ odd
\[
\tilde{z} \overline{\tilde{\alpha}_{n-1}} + \overline{\tilde{\alpha}_{n}} =
e(-nx-y) (z\overline{\alpha_{n-1}}+\overline{\alpha_{n}}).
\]
By Lemma~\ref{lem:tridiag}, we thus obtain that for $n$ even
we have that
\begin{align*}
(\tilde{z} (\widetilde{\mathcal{L}}^{[a,b]}_{\beta,\gamma})^*
- \widetilde{\mathcal{M}}^{[a,b]}_{\beta,\gamma}) U \psi (n) &=
e(-x) z \rho_{n} u_{n+1} \psi(n+1) - \rho_{n-1} u_{n-1} \psi(n-1)\\
& + u_n e((n-1) x +y) (z\alpha_n+\alpha_{n-1}) \psi(n).
\end{align*}
Since $u_{n} = e(-(n-1) x - y) u_{n-1}$ and $u_{n+1} = e(x)u_{n-1}$,
the claimed equality follows for $n$ even.
Similarly for $n$ odd
\begin{align*}
(\tilde{z} (\widetilde{\mathcal{L}}^{[a,b]}_{\beta,\gamma})^*
- \widetilde{\mathcal{M}}^{[a,b]}_{\beta,\gamma}) U \psi (n) &=
- \rho_n u_{n+1} \psi_{n+1} + e(-x) z \rho_{n-1} u_{n-1} \psi_{n-1} \\
& - e(-nx - y) (z \overline{\alpha_{n-1}} + \overline{\alpha_{n}}) u_n \psi(n).
\end{align*}
Since $u_{n+1} = u_n \cdot e(-nx-y)$ and $u_{n} = e((n-1)x + y) u_{n-1}$,
we obtain the claim.
\end{proof}
\begin{proof}[Proof of Proposition~\ref{prop:rotatealphas}]
Since the spectra of $\mathcal{E}^{[a,b]}_{\beta,\gamma}$
and $e(x) \widetilde{\mathcal{E}}^{[a,b]}_{\tilde\beta,\tilde\gamma}$ are
simple, it suffices to show that they are the same.
If $\mathcal{E}^{[a,b]}_{\beta,\gamma} \psi = z \psi$
for $\psi\neq 0$,
we have that $(z (\mathcal{L}^{[a,b]}_{\beta,\gamma})^*
- \mathcal{M}^{[a,b]}_{\beta,\gamma}) \psi = 0$.
Hence, by the previous lemma also that
\[
(\tilde{z} (\widetilde{\mathcal{L}}^{[a,b]}_{\tilde\beta,\tilde\gamma})^*
- \widetilde{\mathcal{M}}^{[a,b]}_{\tilde\beta,\tilde\gamma}) \varphi =0
\]
for $\varphi = U \psi \neq 0$. Hence, we also have that
\[
(z - e(x) \widetilde{\mathcal{E}}^{[a,b]}_{\tilde\beta,\tilde\gamma}) \varphi =0
\]
which implies the claim.
\end{proof}
We will now begin drawing conclusions from Proposition~\ref{prop:rotatealphas}.
For the sake of concreteness, we will only consider
the Verblunsky coefficients given by
\begin{equation}
\alpha_{x,n} = \lambda \mathrm{e}^{2\pi\mathrm{i} (T^n x)_k}
\end{equation}
where $x\in{\mathbb T}^k$, $\lambda\in{\mathbb D}\setminus\{0\}$,
and $T:{\mathbb T}^k\to{\mathbb T}^k$ is the
$k$ dimensional skew-shift defined in \eqref{eq:defkskew}.
For $\theta_1, \theta_2 \in {\mathbb T}$, we denote by
$P_{[\theta_1,\theta_2]}$ the spectral projection
on the arc $\{\mathrm{e}^{2\pi\mathrm{i} t}:\quad t \in [\theta_1, \theta_2] \pmod{1}\}$.
We then have that
\begin{theorem}\label{thm:wegneresti}
Let $\beta_0,\gamma_0 \in\partial{\mathbb D}$ and define
\begin{equation}
\beta_x = \beta_0 \frac{\alpha_{x,-1}}{|\alpha_{x,-1}|},\quad
\gamma_x = \gamma_0 \frac{\alpha_{x,n-1}}{|\alpha_{x,n-1}|}.
\end{equation}
Then
\begin{equation}
\frac{1}{n} \int_{{\mathbb T}^k}\mathrm{tr}\left(P_{[\vartheta_1, \vartheta_2]}
\mathcal{E}^{[0,n-1]}_{x; \beta_x,\gamma_x} \right) dx
= |\theta_2 - \theta_1|.
\end{equation}
\end{theorem}
\begin{proof}
We will show this is true, when only performing the $x_{k-1}$
integral. Let $s = x_{k-1}$. Then changing $s$ amounts to changing
$x$ in Proposition~\ref{prop:rotatealphas}. Hence, the eigenvalues
are given by
\[
\mathrm{e}^{2\pi\mathrm{i} (\theta_1 - s)},\dots, \mathrm{e}^{2\pi\mathrm{i} (\theta_N - s)}
\]
as $s$ varies. This implies the claim.
\end{proof}
It is easy to infer from this that the integrated
density of states is just given by the normalized
Lebesgue measure.
We now come to
\begin{theorem}\label{thm:lyappos}
For $z\in \partial{\mathbb D}$, we have that
\begin{equation}
\gamma(z) = - \frac{1}{2} \log(1 - |\lambda|^2).
\end{equation}
\end{theorem}
\begin{proof}
This can be shown as in Theorem~12.6.2. in \cite{opuc2}.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:main2}]
For $\theta\in{\mathbb T}$, we have
\[
\alpha_{\tilde{x}, n} = \mathrm{e}^{2\pi\mathrm{i} \theta} \alpha_{x, n}
\]
where
\[
\tilde{x}_{\ell} = \begin{cases}
x_{\ell}, &1\leq \ell \leq k-1; \\
x_{k} + \theta,&\ell=k.\end{cases}.
\]
The claim now follows from Theorem~12.6.1.
in \cite{opuc2}.
\end{proof}
\begin{proof}[Proof of Proposition~\ref{prop:rotatealpha}]
If $z \in \sigma_{\mathrm{ess}}(\mathcal{C})$ then there
exists a sequence $\psi_j \in \ell^2({\mathbb N})$ such that
$\|\psi_j\| = 1$, $\psi_j \to 0$ weakly,
and $\|(\mathcal{C} - z)\psi_j\| \to 0$. In particular,
we have for any $N \geq 1$ fixed
\[
\sum_{n=1}^{N} |\psi_j(n)|^2 \to 0.
\]
By Lemma~\ref{lem:rotatealpha} with $x = 2\pi\eta$,
$y = 0$, we obtain that $\varphi_j = U \psi_j$
satisfy $\varphi_j \to 0$ weakly and
\[
\|(\widetilde{\mathcal{C}} - \mathrm{e}^{-2\pi\mathrm{i} \eta} z) \varphi_j\|
\to 0.
\]
Hence, the claim follows.
\end{proof}
\section{Eigenvalue statistics and the proof of Theorem~\ref{thm:main4}}
\label{sec:evstat}
Since we will focus on the case $k=2$, it will be convenient
to introduce the skew-shift $T:{\mathbb T}^2\to{\mathbb T}^2$ by
\begin{equation}
T(x,y) = (x + 2\omega, x + y) \pmod{1}.
\end{equation}
One easily checks that this is equivalent
to \eqref{eq:defkskew} and that
\begin{equation}
T^n(x,y) = (x + 2 n \omega, y + n x+ n(n-1)\omega)\pmod{1}.
\end{equation}
Then our Verblunsky coefficients are given by
\begin{equation}
\alpha_{x,y; n} = \lambda e(y + n x + n(n-1)\omega),
\end{equation}
where we use the abbreviation $e(t) = \mathrm{e}^{2\pi\mathrm{i} t}$.
The main goal of this section is to prove the following
theorem, which will imply Theorem~\ref{thm:main4}.
\begin{theorem}\label{thm:evskew}
Assume $\omega$ satisfies \eqref{eq:conddiop}.
Let $x,y\in{\mathbb T}$ and $\beta,\gamma\in\partial{\mathbb D}$.
There exists $\sigma > 0$ such that for $N$
sufficiently large, there exist
$\theta_1^N, \dots, \theta_N^N$ and $\vartheta^N$
such that
\begin{equation}
\sigma(\mathcal{E}_{x,y; \beta,\gamma}^{[0,N-1]})
= \left\{\mathrm{e}^{2\pi\mathrm{i} \theta_1^N}, \dots, \mathrm{e}^{2\pi\mathrm{i}\theta_N^N}
\right\}
\end{equation}
and
\begin{equation}
\frac{1}{N} \#\left\{n:\quad \|\theta_n^N - \vartheta^N + 2 n \omega\|
> \frac{1}{N^{1 + \sigma}}\right\} \leq \frac{1}{N^{\sigma}}.
\end{equation}
\end{theorem}
In order to see how this implies Theorem~\ref{thm:main4},
we need to introduce some more notation related to the
Laplace functional. Given $N$ points $x_1^N, \dots, x_N^N \in {\mathbb T}$,
we define for $\theta\in{\mathbb T}$
\begin{equation}
\left[-\frac{1}{2},\frac{1}{2}\right) \ni x_n^N(\theta)
= x_n^N - \theta \pmod{1}.
\end{equation}
Then their Laplace functional is defined by
\begin{equation}
\mathfrak{L}_{x^N, N}(f) = \int_{{\mathbb T}} \exp\left(-\sum_{n=1}^{N}
f(N x_n^N(\theta))\right) d\theta
\end{equation}
where $f$ is a continuous, compactly supported, and positive
function. If $\underline{x} = \{ \{x_n^{N}\}_{n=1}^{N} \}_{N=1}^{\infty}$
is a sequence of vectors, we denote
\begin{equation}
\mathfrak{L}_{\underline{x}, N}(f) = \mathfrak{L}_{x^{N}, N}(f).
\end{equation}
Theorem~\ref{thm:main4} follows by applying
(iv) of the next lemma to the sequences
\begin{equation}
\underline{\theta} = \{\{\theta_n^{N}\}_{n=1}^{N}\}_{N=1}^{\infty},\quad
\underline{\vartheta} = \{\{\vartheta^{N} - 2 n\omega\}_{n=1}^{N}\}_{N=1}^{\infty}
\end{equation}
\begin{lemma}
Let $f:{\mathbb R}\to{\mathbb R}$ be a positive, continuous, and
compactly supported function,
$\underline{x} = \{ \{x_n^{N}\}_{n=1}^{N}\}_{N=1}^{\infty}$
and $\underline{y}$ be sequences of vectors in ${\mathbb T}$.
\begin{enumerate}
\item Let $c > 0$ and $A > 1$, then
\begin{equation}
|\{\theta\in{\mathbb T}:\quad\#\{1\leq n\leq N:\quad N x_n^N(\theta)\in
[-c,c]\} \geq A \}| \leq \frac{2c}{N}.
\end{equation}
\item If $\max_{1\leq n\leq N} N \|x_n^{N} - y_n^{N}\| \to 0$
then
\begin{equation}
|\mathfrak{L}_{\underline{x},N}(f) -
\mathfrak{L}_{\underline{y},N}(f)| \to 0.
\end{equation}
\item If
\begin{equation}
\frac{1}{N} \#\{1\leq n \leq N:\quad x_n^{N} \neq y_n^{N}\} \to 0
\end{equation}
then
\begin{equation}
|\mathfrak{L}_{\underline{x},N}(f) -
\mathfrak{L}_{\underline{y},N}(f)| \to 0.
\end{equation}
\item If for every $\varepsilon > 0$
\begin{equation}
\frac{1}{N} \#\{1\leq n \leq N:\quad \|x_n^{N} - y_n^{N}\| \geq \frac{\varepsilon}{N}\} \to 0
\end{equation}
then
\begin{equation}
|\mathfrak{L}_{\underline{x},N}(f) -
\mathfrak{L}_{\underline{y},N}(f)| \to 0.
\end{equation}
\end{enumerate}
\end{lemma}
\begin{proof}[Proof of (i)]
Follows from
\[
\int_{-\frac{1}{2}}^{\frac{1}{2}}
\sum_{n=1}^{N} \chi_{\left[-\frac{c}{N}, \frac{c}{N}\right]}
(x_n^{N}(\theta)) d\theta = 2 c.
\]
and Markov's inequality.
\end{proof}
\begin{proof}[Proof of (ii)]
Let $\varepsilon > 0$.
Since $f$ is compactly supported, we have $\mathrm{supp}(f)
\subseteq [-c,c]$. Let $A = \lceil \frac{2c}{10 \varepsilon} \rceil$.
By (i), there exists a set $I \subseteq {\mathbb T}$ such that
for $\theta \in I$
\[
\#\{1\leq n \leq N:\quad N x_n(\theta) \in [-c,c]
\text{ or } N y_n(\theta) \in [-c,c] \} \leq A
\]
and $|{\mathbb T} \setminus I| \leq \frac{\varepsilon}{2}$.
By assumption, $f$ is uniformly continuous, so there
exists a $\delta > 0$ such that $|f(x) - f(y)| \leq \frac{\varepsilon}{2 A}$
for $|x-y| < \delta$. Choose $N$ so large that
\[
\max_{1\leq n \leq N} N \|x_n^{N} - y_n^{N}\| < \delta.
\]
Then we clearly have that $|N x_n(\theta) - N y_n(\theta)| < \delta$,
and thus that for $\theta \in I$.
\[
\left|\sum_{n=0}^{N-1} f(N x_n(\theta))
- \sum_{n=0}^{N-1} f(N y_n(\theta))\right| < \frac{\varepsilon}{2}.
\]
The claim follows.
\end{proof}
\begin{proof}[Proof of (iii)]
(ii) follows from the set of $\theta$ for which
\[
\sum_{n=1}^{N} f(N y_n^N(\theta)) \neq
\sum_{n=1}^{N} f(N x_n^N(\theta))
\]
having vanishing measure as $N\to\infty$.
\end{proof}
\begin{proof}[Proof of (iv)]
By assumption, there exists $\varepsilon_N \to 0$ such that
\[
\frac{1}{N} \#\{1\leq n\leq N:\quad \|x_n^{N} - y_n^{N} \|
\geq \frac{\varepsilon_N}{N} \} \to 0.
\]
Define
\[
\tilde{x}_n^{N} = \begin{cases} y_n^{N},& \|x_n^{N} - y_n^{N}\|
\geq \frac{\varepsilon_N}{N};\\
x_n^N, & \text{otherwise}.\end{cases}
\]
Then $x^N$ and $\tilde{x}^N$ satisfy the assumptions of (i)
and $\tilde{x}^N$ and $y^N$ the ones of (ii). The claim follows.
\end{proof}
We now begin the proof of Theorem~\ref{thm:evskew}.
\eqref{eq:conddiop} implies that there exists
some $c > 0$ such that
\begin{equation}\label{eq:conddiop2}
\|q \omega\| \geq \frac{c}{q^{\tau}}
\end{equation}
for all positive integers $q$. The following
theorem will be essential to our proof and
proven only in the next section.
\begin{theorem}\label{thm:existtest}
There is a constant $\sigma \in (0,1)$.
Let $\eta \geq 1$, $N$ sufficiently large,
$\beta,\gamma\in\partial{\mathbb D}$
and $x,y\in{\mathbb T}$. There exists a normalized
$\psi \in \ell^2(\{0,\dots,N-1\})$ such that
$\psi(n) = 0$ for $n \geq N^{\sigma}$, $n =0,1$
and $z = \mathrm{e}^{2\pi\mathrm{i} \vartheta}$ such that
\begin{equation}
\|(\mathcal{E}^{[0,N-1]}_{x,y;\beta,\gamma} - z) \psi\|
\leq \left(\frac{1}{N}\right)^{\eta}.
\end{equation}
\end{theorem}
Define $\vartheta_k = \vartheta - 2 \omega k$ and
$z_k = e(\vartheta_k)$.
\begin{lemma}
Let $\varepsilon > 0$.
If \eqref{eq:conddiop2} holds, then for
$N$ large enough and $k \neq \tilde{k} \in \{0, \dots, N-1\}$
we have
\begin{equation}
|z_k - z_{\tilde{k}}| \geq \frac{1}{N^{\tau + \varepsilon}}.
\end{equation}
\end{lemma}
\begin{proof} Clear. \end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:evskew}]
We let $\eta = \tau + 2\varepsilon$ in Theorem~\ref{thm:existtest}.
With $u_{n}$ the appropriate factors as given
in Lemma~\ref{lem:rotatealpha}, we define the test
functions
\[
\psi_k(n) = \begin{cases} u_{n} \psi(n - k), &k\leq n\leq k+N^{\sigma}\\
0, &\text{otherwise}. \end{cases}
\]
We then have for $0 \leq k \leq N - N^{\sigma}$ that
\[
\|(\mathcal{E}_{x,y;\beta,\gamma}^{[0,N-1]}-z_k) \psi_k\|
\leq \frac{1}{N^{\eta}}.
\]
Hence, there is some eigenvalue $\mathrm{e}^{2\pi\mathrm{i}\theta_{\ell_k}}$
such that
\[
\|\theta_{\ell_{k}} - \vartheta_k\| \leq \frac{1}{N^{\eta}}.
\]
By the previous lemma, we must have $\ell_k \neq \ell_{\tilde{k}}$
for $k \neq \tilde{k}$. The claim then follows upon
reordering the $\theta_{\ell}$.
\end{proof}
\section{Proof of Theorem~\ref{thm:existtest}}
\label{sec:existtest}
Let $L = \lfloor \frac{1}{3} N^{\sigma}\rfloor$.
If we show that for every $(x,y)\in{\mathbb T}^2$, there exists
a normalized vector $\psi$ and $z\in\partial{\mathbb D}$
such that
\begin{equation}
\|(\mathcal{E}^{[-L,L]}_{x,y; \beta,\gamma} - z) \psi\|
\leq \frac{1}{N^C}
\end{equation}
then Theorem~\ref{thm:existtest} follows. We will
show this modified claim, since it is notationally
somewhat simpler to deal with.
Since $\mathcal{E}^{[-L,L]}_{x,y; \beta,\gamma}$ has $2 L + 1$ eigenvalues,
there exists $z\in{\mathbb D}$ and $\|\psi\| = 1$ such that
\begin{equation}
\mathcal{E}^{[-L,L]}_{x,y; \beta,\gamma} \psi = z \psi,
\quad |\psi(0)|^2 \geq \frac{1}{2L+1}.
\end{equation}
We will prove in the following section
\begin{theorem}\label{thm:greendecays}
There exists $\eta > 0$ such that for every $C \geq 1$,
we have for $L$ large enough and $M = \lfloor L^{\eta} \rfloor$
that there exist
\begin{equation}
-\frac{2}{3} L \leq k_- \leq -\frac{1}{3} L,\quad
\frac{1}{3} L \leq k_+ \leq \frac{2}{3} L
\end{equation}
such that for
\begin{equation}
k \in \{k_- - C M, \dots, k_- + C M\}
\cup \{k_+ - CM, \dots, k_+ + C M\}
\end{equation}
we have that there
exist $\beta, \gamma \in \partial{\mathbb D}$ such that
for $|k - \ell| \leq \frac{M}{2}$ we have
\begin{equation}
|G^{[k-M, k+M]}_{x,y;\beta,\gamma}(z; \ell, k-M)|,\
|G^{[k-M, k+M]}_{x,y;\beta,\gamma}(z; \ell, k+M)| \leq \frac{1}{M}.
\end{equation}
\end{theorem}
Define
\begin{equation}
\mathcal{K}_{t} =
\{k_- - t M, \dots, k_- + t M\}
\cup \{k_+ - t M, \dots, k_+ + t M\}.
\end{equation}
Using Lemma~\ref{lem:green2sol} combined
with the estimate from the previous theorem, we
can conclude for $k \in K_{C}$ and $|\ell -k| \leq \frac{M}{2}$
that
\begin{equation}
|\psi(\ell)| \leq \frac{4}{M},
\end{equation}
where we used the trivial estimate $|\psi(n)|\leq 1$
We can iterate this to obtain for $s = 1,\dots, C$ that
for $k \in K_{C-s+1}$ and $|\ell -k| \leq \frac{M}{2}$
\begin{equation}
|\psi(\ell)| \leq \left(\frac{4}{M}\right)^s.
\end{equation}
In particular, we obtain that
\begin{equation}
|\psi(k_-)|, |\psi(k_+)| \leq \left(\frac{4}{M}\right)^C.
\end{equation}
Define a test function $\varphi$ by
\begin{equation}
\varphi(n) = \begin{cases} \psi(n), & k_- \leq n \leq k_+; \\
0,&\text{otherwise}.\end{cases}
\end{equation}
We have that
\begin{equation}
\|(\mathcal{E}^{[-L,L]}_{x,y;\beta,\gamma} - z) \varphi\|
\leq \left(\frac{8}{L}\right)^{C \eta}
\end{equation}
and thus Theorem~\ref{thm:existtest} follows.
\section{Decay of the Green's function:
Proof of Theorem~\ref{thm:greendecays}}
\label{sec:greendecays}
Theorem~\ref{thm:uegreen} states that
$m = L(z) > 0$ implies that for $N$ large enough
there exists a set $B_N \subseteq{\mathbb T}^2$ with
\begin{enumerate}
\item $|B_N|\to 0$.
\item For $(x,y) \in {\mathbb T}^2\setminus B_N$, there exists
\begin{equation}
\beta \in \left\{-\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|},
\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|}\right\},\quad
\gamma \in \left\{\frac{\alpha_N}{|\alpha_N|},
-\frac{\alpha_N}{|\alpha_N|}\right\}
\end{equation}
such that for $|k| \leq \frac{N}{2}$, we have
\begin{equation}
|G^{[-N,N]}_{x,y;\beta,\gamma}(z; -N, k)|
\leq \mathrm{e}^{-\frac{m}{4} |k+N|},
\quad
|G^{[-N,N]}_{x,y;\beta,\gamma}(z; N, k)|
\leq \mathrm{e}^{-\frac{m}{4} |N-k|}.
\end{equation}
\end{enumerate}
We will first need the following lemma.
\begin{lemma}
Let $\sigma > 0$, there exists a constant $C > 1$ such
that for $N \geq 1$, there exists a set $B^2_N$ such
that
\begin{equation}
|B_N^2| \leq \frac{C}{N^{\sigma}}
\end{equation}
and for
\begin{equation}
\beta \in \left\{-\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|},
\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|}\right\},\quad
\gamma \in \left\{\frac{\alpha_N}{|\alpha_N|},
-\frac{\alpha_N}{|\alpha_N|}\right\}
\end{equation}
and $x,y\in{\mathbb T}^2\setminus B_N^{2}$ we have
\begin{equation}
\left\|\left(z \left(
\mathcal{L}^{[-N,N]}_{x,y;\beta,\gamma}\right)^{*} -
\mathcal{M}^{[-N,N]}_{x,y;\beta,\gamma}\right)^{-1}
\right\|
\leq N^{1 + \sigma}
\end{equation}
\end{lemma}
\begin{proof}
This is a consequence of Theorem~\ref{thm:wegneresti}.
\end{proof}
In summary, we have extracted the following statement
\begin{proposition}
Let $\sigma > 0$. For $N \geq 1$ large enough, there
exists $\Omega_N \subseteq{\mathbb T}^2$ such that
\begin{equation}
\lim_{N\to\infty} |{\mathbb T}^2 \setminus\Omega_N| = 0.
\end{equation}
For each $(x,y) \in \Omega_N$ and
\begin{equation}
|\tilde{x} - x| \leq \frac{1}{N^{2(1+2\sigma)}},\quad
|\tilde{y} - y| \leq \frac{1}{N^{1+2\sigma}},
\end{equation}
we have that there exists
\begin{equation}
\beta \in \left\{-\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|},
\frac{\alpha_{-N-1}}{|\alpha_{-N-1}|}\right\},\quad
\gamma \in \left\{\frac{\alpha_N}{|\alpha_N|},
-\frac{\alpha_N}{|\alpha_N|}\right\}
\end{equation}
such that for $|k| \leq \frac{N}{2}$, we have
\begin{equation}
|G^{[-N,N]}_{\tilde x, \tilde y;\beta,\gamma}(z; -N, k)| \leq \frac{1}{N},
\quad
|G^{[-N,N]}_{\tilde x, \tilde y;\beta,\gamma}(z; N, k)| \leq \frac{1}{N}.
\end{equation}
\end{proposition}
\begin{proof}
A computation shows that
\[
\|\mathcal{L}^{[-N,N]}_{x,y;\beta,\gamma} -
\mathcal{L}^{[-N,N]}_{\tilde{x},\tilde{y};\beta,\gamma}\|
\lesssim N |x - \tilde{x}| + |y + \tilde{y}|
\leq \frac{1}{N^{\frac{\sigma}{2}}}
\]
for $N$ large enough
and a similar result for
$\mathcal{M}^{[-N,N]}_{\tilde{x},\tilde{y};\beta,\gamma}$. The result now
follows from
\[
B^{-1} - A^{-1} = B^{-1} (A - B) A^{-1}
\]
and some computations.
\end{proof}
Let $X_N = \lceil N^{2 (1 + 2\sigma)} \rceil$,
$Y_N = \lceil N^{1+2\sigma} \rceil$. We partition
${\mathbb T}^2$ into $X_N \cdot Y_N \lesssim N^{3(1+2\sigma)}$
boxes of side length $\frac{1}{X_N}$ and $\frac{1}{Y_N}$.
We call a box $I_{\ell}$ bad if
\begin{equation}
\Omega_N \cap I_{\ell} = \emptyset
\end{equation}
and good otherwise. We note that if $(x,y)$ is in a good
box, then for $|k|\leq\frac{N}{2}$
\begin{equation}
|G^{[-N,N]}_{x,y; \beta, \gamma}(z; k, \pm N)| \leq \frac{1}{N}
\end{equation}
for some $\beta, \gamma\in\partial{\mathbb D}$.
We now given an upper bound on the number of iterates
of $T^{j}(x,y)$ that land in any bad box.
We will show the following theorem in Appendix~\ref{sec:dynskew}.
For $\varepsilon, \delta >0$, denote by $B_{\varepsilon,\delta}
\subseteq{\mathbb T}^2$
the set
\begin{equation}
B_{\varepsilon,\delta} = \{(x,y)\in{\mathbb T}^2:\quad
\|x\|\leq \varepsilon,\ \|y\|\leq\delta\}.
\end{equation}
\begin{theorem}\label{thm:skewergodic2}
Assume \eqref{eq:conddiop} and let
$\delta > 0$, $\varepsilon > 0$, $N \geq 1$. There exists $L_0 = L_0(\sigma,\omega) \geq 1$
such that for any $x,y\in{\mathbb T}$ there exists $0 \leq \ell_0 \leq N$
such that for $L \geq L_0 \delta^{-4} \varepsilon^{-9}$
\begin{equation}
\#\{0 \leq \ell \leq \frac{L_0}{N}:\quad
T^{\ell N}(x,y) \in B_{\varepsilon, \delta}\} \leq 10 \varepsilon \delta \frac{L}{N}.
\end{equation}
\end{theorem}
We now obtain that for $L \geq N^{15}$ and $N$ large enough, we have
for some $0 \leq \ell_0 \leq N-1$
\begin{equation}
\#\{\lfloor\frac{1}{3N } L+\ell_0\rfloor \leq \ell \leq \frac{2}{3N} L:
\quad T^{\ell N}(x,y)\text{ in fixed
bad box}\} \leq \frac{10 L/N}{N^{3 (1 + \sigma)}}.
\end{equation}
Since
\begin{equation}
\#\{\text{bad boxes}\} \leq \delta_N N^{3(1+\sigma)}
\end{equation}
with $\delta_N \to 0$ as $N \to \infty$, we obtain for $L \geq N^{15}$ that
\begin{equation}
\#\{\lfloor\frac{1}{3N } L+\ell_0\rfloor \leq \ell \leq \frac{2}{3N} L:
\quad T^{\ell N}(x,y)\text{ in some bad box}\} \leq \delta_N \frac{L}{N}
\end{equation}
for $\delta_N\to 0$ as $N \to \infty$.
\begin{proof}[Proof of Theorem~\ref{thm:greendecays}]
We just give the argument for $k_+$.
Choose $\delta_N \leq \frac{1}{10 C}$.
Now divide $\left[\lfloor\frac{1}{3N } L+\ell_0\rfloor,
\frac{2}{3N} L\right]$ into segments of length $3 C$.
Then at most $\delta_N \frac{L}{N}$ of them can
contain an iterate that lands in a bad box,
but there are $\frac{1}{C} \frac{L}{3 N} = 10 \delta_N
\frac{L}{3N}$ many of them. Hence, we must have at
least one, where our conclusion holds.
\end{proof}
|
\section{Introduction}
\label{s1}
One of fundamental issues in nuclear physics is the equation of
state (EOS) of isospin asymmetric nuclear matter, which plays a
central role in understanding not only the structure of radioactive
nuclei, the reaction dynamics induced by rare isotopes, and the
liquid-gas phase transition in asymmetric nuclear matter, but also
many critical issues in astrophysics
\cite{LiBA98,danie02,lattimer04,baran05,steiner05,chen07,li08}. For
symmetric nuclear matter with equal fractions of neutrons and
protons, its EOS is relatively well-determined from analyses of the
giant monopole resonances of finite nuclei \cite{youngblood99,LiT07}
as well as collective flows \cite{danie02} and subthreshold kaon
production \cite{aic85, fuchs06} in relativistic nucleus-nucleus
collisions. On the other hand, the EOS of asymmetric nuclear matter,
especially the density dependence of the nuclear symmetry energy
$E_{\mathrm{sym}}(\rho )$, is poorly known. During the last decade,
significant progress has been made both experimentally and
theoretically on constraining the behavior of the symmetry energy
around and below normal nuclear matter
density~\cite{Che05,Tsa09,Cen09,Nat10} (See, e.g., Refs.
\cite{XuC10,Che10,Tsa11,Che11a,New11} for review of recent progress)
while its super-normal density behavior remains elusive and largely
controversial~\cite{Xia09,Fen10,Rus11,XuC10b}. Theoretically, all
many-body theory calculations to date have demonstrated that the
nuclear symmetry energy essentially characterizes the isospin
dependent part of the EOS of asymmetric nuclear matter and the
higher-order terms in isospin
asymmetry are unimportant, at least for densities up to moderate values \cit
{li08}, leading to the well-known empirical parabolic law.
When the empirical parabolic law itself provides a good approximation to the
EOS of asymmetric nuclear matter and thus allows one to extract the symmetry
energy from the energy difference between pure neutron matter and symmetric
nuclear matter, it may cause large errors when it is applied to determine
some physical quantities under special conditions. For example, the
higher-order terms in isospin asymmetry presented in the EOS of asymmetric
nuclear matter at supra-normal densities can significantly modify the proton
fraction in $\beta $-equilibrium neutron-star matter and the critical
density for the direct Urca process which can lead to faster cooling of
neutron stars \cite{Zha01,steiner08}. In addition, recent studies \cite{xu09}
indicate that the higher-order terms in isospin asymmetry are very important
for determining the transition density and pressure at the inner edge
separating the liquid core from the solid crust of neutron stars where the
matter is extremely neutron-rich. Furthermore, the higher-order effects on
the incompressibility of asymmetric nuclear matter have also been studied
recently \cite{Che09}. These studies about the higher-order effects are
essentially performed within the nonrelativistic models since the analytical
expressions of the higher-order terms in isospin asymmetry, e.g., the
nuclear matter fourth-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$,
can be relatively easily obtained in such nonrelativistic models. It is thus
interesting to see if the same conclusion can be obtained within the
relativistic models.
One of very popular relativistic models is the relativistic mean field (RMF)
model which is generally based on effective interaction Lagrangians
involving nucleon and meson fields \cite{Wal74}. As a phenomenological
approach, the RMF model has achieved great success during the last decades
in describing many nuclear phenomena \cite{Ser86,Men06}. Although the full
expressions have been usually used in realistic RMF model calculations,
nevertheless, it will be instructive to see separately the effects of the
higher-order terms in isospin asymmetry, e.g., the nuclear matter
fourth-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$, within the RMF
model. The main motivation of the present work is to derive the nuclear
matter fourth-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$ within the
nonlinear RMF model and then explore the higher-order $E_{\mathrm{{sym},4
}(\rho )$ corrections to the widely used empirical parabolic law for the
isospin asymmetric nuclear matter. Based on two accurately calibrated
interactions, our results indicate that the $E_{\mathrm{{sym},4}}(\rho )$
may have significant influence on the properties of isospin asymmetric
nuclear matter, the proton fraction $x_{p}$ in $\beta $-stable neutron star
matter and the core-crust transition density $\rho _{t}$ and pressure $P_{t}$
in neutron stars, confirming the previous nonrelativistic calculations.
The paper is organized as follows. In Section \ref{s2}, we briefly discuss
the bulk characteristic parameters of asymmetric nuclear matter. The model
used in the present paper and the analytical expression of the $4$th-order
symmetry energy will also given in this section. The results and discussions
are then presented in Section \ref{s3}. Finally, a summary is given in
Section \ref{s4}.
\section{Theoretical formulism}
\label{s2}
\subsection{Characteristic parameters of asymmetric nuclear matter}
The EOS of isospin asymmetric nuclear matter, defined by its binding energy
per nucleon, can be expanded to $4$nd-order in isospin asymmetry $\delta $ a
\begin{equation}
E(\rho ,\delta )=E_{0}(\rho )+E_{\mathrm{sym}}(\rho )\delta ^{2}+E_{\mathrm{
sym},4}}(\rho )\delta ^{4}+\mathcal{O}(\delta ^{6}), \label{EOSANM}
\end{equation
where $\rho =\rho _{n}+\rho _{p}$ is the baryon density with $\rho _{n}$ and
$\rho _{p}$ denoting the neutron and proton densities, respectively; $\delta
=(\rho _{n}-\rho _{p})/(\rho _{p}+\rho _{n})$ is the isospin asymmetry;
E_{0}(\rho )=E(\rho ,\delta =0)$ is the binding energy per nucleon in
symmetric nuclear matter; the nuclear matter symmetry energy $E_{\mathrm{sym
}(\rho )$ and the $4$th-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$
are expressed, respectively, as
\begin{eqnarray}
E_{\mathrm{sym}}(\rho ) &=&\left. \frac{1}{2!}\frac{\partial ^{2}E(\rho
,\delta )}{\partial \delta ^{2}}\right\vert _{\delta =0},~~ \label{Esym} \\
E_{\mathrm{{sym},4}}(\rho ) &=&\left. \frac{1}{4!}\frac{\partial ^{4}E(\rho
,\delta )}{\partial \delta ^{4}}\right\vert _{\delta =0}. \label{Esym4}
\end{eqnarray
In Eq. (\ref{EOSANM}), the absence of odd-order terms in $\delta $ is due to
the exchange symmetry between protons and neutrons in nuclear matter when
one neglects the Coulomb interaction and assumes the charge symmetry of
nuclear forces. The higher-order (including the $4$th-order) coefficients in
$\delta $ are usually very small. For example, the magnitude of the $\delta
^{4}$ term at normal nuclear matter density $\rho _{0}$ is estimated to be
less than $1$ MeV in microscopic many-body approaches \cite{sie70,sj74,lag81}
and also in phenomenological nonrelativistic models \cite{Che09} as well as
relativistic models as will be shown in this work. Neglecting the
contribution from higher-order terms in Eq. (\ref{EOSANM}) leads to the
well-known empirical parabolic law, i.e., $E(\rho ,\delta )\simeq E_{0}(\rho
)+E_{\mathrm{sym}}(\rho )\delta ^{2}$ for the EOS of asymmetric nuclear
matter and the symmetry energy $E_{\mathrm{sym}}(\rho )$ can thus be
extracted from $E_{\mathrm{sym}}(\rho )\simeq E(\rho ,\delta =1)-E(\rho
,\delta =0)$.
Around normal nuclear matter density $\rho _{0}$, the $E_{0}(\rho )$ can be
expanded, e.g., up to $4$th-order in density, as,
\begin{equation}
E_{0}(\rho )=E_{0}(\rho _{0})+\frac{K_{0}}{2!}\chi ^{2}+\frac{J_{0}}{3!}\chi
^{3}+\frac{I_{0}}{4!}\chi ^{4}+\mathcal{O}(\chi ^{5}), \label{DenExp0}
\end{equation
where $\chi $ is a dimensionless variable characterizing the deviations of
the density from normal nuclear matter density $\rho _{0}$ and it is
conventionally defined as
\begin{equation}
\chi =\frac{\rho -\rho _{0}}{3\rho _{0}}. \label{DefTh}
\end{equation
The first term $E_{0}(\rho _{0})$ on the right-hand-side (r.h.s) of Eq. (\re
{DenExp0}) is the binding energy per nucleon in symmetric nuclear matter at
normal nuclear matter density $\rho _{0}$ and the coefficients of other
terms are,
\begin{eqnarray}
K_{0} &=&\left. 9\rho _{0}^{2}\frac{\partial ^{2}E_{0}(\rho )}{\partial \rho
^{2}}\right\vert _{\rho =\rho _{0}},~~ \label{K0} \\
J_{0} &=&\left. 27\rho _{0}^{3}\frac{\partial ^{3}E_{0}(\rho )}{\partial
\rho ^{3}}\right\vert _{\rho =\rho _{0}}, \label{J0} \\
I_{0} &=&\left. 81\rho _{0}^{4}\frac{\partial ^{2}E_{0}(\rho )}{\partial
\rho ^{4}}\right\vert _{\rho =\rho _{0}}. \label{I0}
\end{eqnarray
The linear $\chi $ term on the r.h.s of Eq. (\ref{DenExp0}) vanishes
according to the definition of the saturation density $\rho _{0}$. The
coefficient $K_{0}$ is the well-known incompressibility coefficient of
symmetric nuclear matter and it characterizes the curvature of $E_{0}(\rho
_{0})$ at $\rho _{0}$. The coefficients $J_{0}$ and $I_{0}$ are the $3
rd-order and $4$th-order incompressibility coefficients of symmetric nuclear
matter \cite{Che09}, respectively.
Similarly, around normal nuclear matter density $\rho _{0}$, the symmetry
energy $E_{\mathrm{sym}}(\rho )$ and the $4$th-order symmetry energy $E_
\mathrm{{sym},4}}(\rho )$ can be expanded, e.g., up to $4$th-order in $\chi
, as
\begin{eqnarray}
E_{\mathrm{sym}}(\rho ) &=&E_{\mathrm{sym}}(\rho _{0})+L\chi +\frac{K_
\mathrm{sym}}}{2!}\chi ^{2} \notag \\
&&+\frac{J_{\mathrm{sym}}}{3!}\chi ^{3}+\frac{I_{\mathrm{sym}}}{4!}\chi ^{4}
\mathcal{O}(\chi ^{5}), \label{DenExp2}
\end{eqnarray
and
\begin{eqnarray}
E_{\mathrm{{sym},4}}(\rho ) &=&E_{\mathrm{{sym},4}}(\rho _{0})+L_{\mathrm{
sym},4}}\chi +\frac{K_{\mathrm{{sym},4}}}{2!}\chi ^{2} \notag \\
&&+\frac{J_{\mathrm{{sym},4}}}{3!}\chi ^{3}+\frac{I_{\mathrm{{sym},4}}}{4!
\chi ^{4}+\mathcal{O}(\chi ^{5}), \label{DenExp4}
\end{eqnarray
respectively, where the $L$, $K_{\mathrm{sym}}$, $J_{\mathrm{sym}}$, $I_
\mathrm{sym}}$ and $L_{\mathrm{{sym},4}}$, $K_{\mathrm{{sym},4}}$, $J_
\mathrm{{sym},4}}$, $I_{\mathrm{{sym},4}}$ are the slope parameter,
curvature parameter, $3$rd-order and $4$th-order density coefficients of the
$E_{\mathrm{sym}}(\rho )$ and $E_{\mathrm{{sym},4}}(\rho )$ at $\rho _{0}$,
respectively, whose definitions are similar to Eq. (\ref{K0}) - Eq. (\ref{I0
). In general, these characteristic parameters can be written as,
\begin{equation}
W_{ij}=\left. (3\rho _{0})^{j}\frac{\partial ^{j}E_{\mathrm{sym},2i}(\rho )}
\partial \rho ^{j}}\right\vert _{\rho =\rho _{0}},~~i,j=1,2,\cdots
\label{CoExp}
\end{equation
for example, $W_{11}=L$, $W_{12}=K_{\mathrm{sym}}$, $W_{23}=J_{\mathrm{{sym
,4}}$, $W_{24}=I_{\mathrm{{sym},4}}$, and so on.
In the above Taylor's expansions, we have kept all terms up to $4$th-order
in $\delta $ or $\chi $. The $14$ characteristic parameters, namely,
E_{0}(\rho _{0})$, $K_{0}$, $J_{0}$, $I_{0}$, $E_{\mathrm{sym}}(\rho _{0})$,
$L$, $K_{\mathrm{sym}}$, $J_{\mathrm{sym}}$, $I_{\mathrm{sym}}$, $E_{\mathrm
{sym},4}}(\rho _{0})$, $L_{\mathrm{{sym},4}}$, $K_{\mathrm{{sym},4}}$, $J_
\mathrm{{sym},4}}$ and $I_{\mathrm{{sym},4}}$ are well-defined, and they
characterize the EOS of an asymmetric nuclear matter and its density
dependence at normal nuclear matter density $\rho _{0}$. Among these
parameters, $E_{0}(\rho _{0})$, $K_{0}$, $E_{\mathrm{sym}}(\rho _{0})$, $L$
and $K_{\mathrm{sym}}$ have been extensively studied in the literature and
significant progress has been made over past few decades \cite{li08}.
The incompressibility of asymmetric nuclear matter is an important quantity
to characterize its EOS. Conventionally, the incompressibility coefficient
is defined at the saturation density where the pressure $P(\rho ,\delta
)=\rho ^{2}{\partial E(\rho ,\delta )}/{\partial \rho }=0$, and it is called
the isobaric incompressibility coefficient \cite{prak85} given b
\begin{equation}
K_{\mathrm{sat}}(\delta )=\left. 9\rho _{\mathrm{sat}}^{2}\frac{\partial
^{2}E(\rho ,\delta )}{\partial \rho ^{2}}\right\vert _{\rho =\rho _{\mathrm
sat}}}. \label{isoincom}
\end{equation
The isobaric incompressibility coefficient $K_{\mathrm{sat}}(\delta )$ thus
only depends on the isospin asymmetry $\delta $. One can show that up to $4
th-order in $\delta $, the $K_{\mathrm{sat}}(\delta )$ can be expressed as
\cite{Che09}
\begin{equation}
K_{\mathrm{sat}}(\delta )=K_{0}+K_{\mathrm{sat,2}}\delta ^{2}+K_{\mathrm
sat,4}}\delta ^{4}+\mathcal{O}(\delta ^{6}), \label{SatIncom}
\end{equation
wit
\begin{align}
K_{\mathrm{{sat},2}}=& K_{\mathrm{sym}}-6L-\frac{J_{0}L}{K_{0}},
\label{SatIncom2} \\
K_{\mathrm{sat,4}}=& K_{\mathrm{sym,4}}-6L_{\mathrm{sym,4}}-\frac{J_{0}L_
\mathrm{sym,4}}}{K_{0}}+\frac{9L^{2}}{K_{0}}-\frac{J_{\mathrm{sym}}L}{K_{0}}
\notag \\
& +\frac{I_{0}L^{2}}{2K_{0}^{2}}+\frac{J_{0}K_{\mathrm{sym}}L}{K_{0}^{2}}
\frac{3J_{0}L^{2}}{K_{0}^{2}}-\frac{J_{0}^{2}L^{2}}{2K_{0}^{3}}.
\label{Ksat4}
\end{align
If we use the parabolic approximation for the EOS of symmetric nuclear
matter, i.e., $E_{0}(\rho )=E_{0}(\rho _{0})+\frac{1}{2}K_{0}^{2}\chi
\mathcal{O}(\chi ^{3}),$ then the $K_{\mathrm{{sat},2}}$ parameter is
reduced t
\begin{equation}
K_{\mathrm{asy}}=K_{\mathrm{sym}}-6L, \label{DefKasy}
\end{equation
and this expression has been extensively used in the literature to
characterize the isospin dependence of the incompressibility of asymmetric
nuclear matter in the literature \cite{baran05,Che05,Cen09,LiT07,lopez88}.
\subsection{The $4$th-order symmetry energy in the nonlinear RMF model}
In the present work, we use the interacting Lagrangian density of the
nonlinear RMF model supplemented with couplings between the isoscalar and
the isovector mesons \cite{Mul96,Hor01,Tod05}, i.e.,
\begin{align}
& \mathcal{L}(\psi ,\sigma ,\omega _{\mu },\vec{\mkern1mu\rho }_{\mu })=\bar
\psi}\left[ \gamma _{\mu }(i\partial ^{\mu }-g_{\omega }\omega ^{\mu
})-(M-g_{\sigma }\sigma )\right] \psi \null \notag \\
& +\frac{1}{2}\left( \partial _{\mu }\sigma \partial ^{\mu }\sigma
-m_{\sigma }^{2}\sigma ^{2}\right) -\frac{1}{3}b_{\sigma }M(g_{\sigma
}\sigma )^{3} \notag \\
& -\frac{1}{4}c_{\sigma }(g_{\sigma }\sigma )^{4}\null+\frac{1}{2}m_{\omega
}^{2}\omega _{\mu }\omega ^{\mu }-\frac{1}{4}F_{\mu \nu }F^{\mu \nu } \notag
\\
& +\frac{1}{4}c_{\omega }(g_{\omega }^{2}\omega _{\mu }\omega ^{\mu })^{2
\null+\frac{1}{2}m_{\rho }^{2}\vec{\mkern1mu\rho }_{\mu }\cdot \vec{\mker
1mu\rho }^{\mu } \notag \\
& -\frac{1}{4}\vec{\mkern1muG}_{\mu \nu }\cdot \vec{\mkern1muG}^{\mu \nu
}-g_{\rho }\vec{\mkern1mu\rho }_{\mu }\cdot \bar{\psi}\gamma ^{\mu }\vec
\mkern1mu\tau }\psi \null \notag \\
& +\frac{1}{2}g_{\rho }^{2}\vec{\mkern1mu\rho }_{\mu }\cdot \vec{\mker
1mu\rho }^{\mu }\left[ \Lambda _{S}g_{\sigma }^{2}\sigma ^{2}+\Lambda
_{V}g_{\omega }^{2}\omega _{\mu }\omega ^{\mu }\right] \label{NonRMF}
\end{align
where $F_{\mu \nu }\equiv \partial _{\mu }\omega _{\nu }-\partial _{\nu
}\omega _{\mu }$ and$~\vec{\mkern1muG}_{\mu \nu }\equiv \partial _{\mu }\vec
\mkern1mu\rho }_{\nu }-\partial _{\nu }\vec{\mkern1mu\rho }_{\mu }$ are
strength tensors for $\omega $ field and $\rho $ field, respectively. $\psi
, $\sigma $, $\omega _{\mu }$, $\vec{\mkern1mu\rho }_{\mu }$ are nucleon
field, isoscalar-scalar field, isoscalar-vector field and isovector-vector
field, respectively, and the arrows denote the vector in isospin space. The
\Lambda _{S}$ and $\Lambda _{V}$ represent coupling constants between the
isovector $\rho $ meson and the isoscalar $\sigma $ and $\omega $ mesons,
respectively, which are important for the description of the density
dependence of the symmetry energy. In addition, $M$ is the nucleon mass and
m_{\sigma }$, $m_{\omega }$, $m_{\rho }$ are masses of mesons.
In the mean field approximation, after neglecting effects of fluctuation and
correlation, meson fields are replaced by their expectation values, i.e.,
\bar{\sigma}\rightarrow \sigma $, $\bar{\omega}_{0}\rightarrow \omega _{\mu
} $, $\bar{\rho}_{0}^{(3)}\rightarrow \vec{\mkern1mu\rho }_{\mu }$, where
subscript \textquotedblleft $0$" indicates zeroth component of the
four-vector, superscript \textquotedblleft ($3$)" indicates third component
of the isospin, Furthermore, we also use in this work the non-sea
approximation which neglects the effect due to negative energy states in the
Dirac sea. The mean field equations are then expressed as
\begin{align}
m_{\sigma }^{2}\bar{\sigma}=& g_{\sigma }\left[ \rho _{S}-b_{\sigma }M\left(
g_{\sigma }\bar{\sigma}\right) ^{2}-c_{\sigma }\left( g_{\sigma }\bar{\sigma
\right) ^{3}+\left( g_{\rho }\bar{\rho}_{0}^{(3)}\right) ^{2}\Lambda
_{S}g_{\sigma }\bar{\sigma}\right] \\
m_{\omega }^{2}\bar{\omega}_{0}=& g_{\omega }\left[ \rho -c_{\omega }\left(
g_{\omega }\bar{\omega}_{0}\right) ^{3}-\Lambda _{V}g_{\omega }\bar{\omega
_{0}\left( g_{\rho }\bar{\rho}_{0}^{(3)}\right) ^{2}\right] \\
m_{\rho }^{2}\bar{\rho}_{0}^{(3)}=& g_{\rho }\left[ \rho _{p}-\rho
_{n}-\Lambda _{S}g_{\rho }\bar{\rho}_{0}^{(3)}\left( g_{\sigma }\bar{\sigma
\right) ^{2}-\Lambda _{V}g_{\rho }\bar{\rho}_{0}^{(3)}\left( g_{\omega }\bar
\omega}_{0}\right) ^{2}\right]
\end{align
where
\begin{equation}
\rho =\langle \bar{\psi}\gamma ^{0}\psi \rangle =\rho _{n}+\rho _{p},~~\rho
_{S}=\langle \bar{\psi}\psi \rangle =\rho _{Sn}+\rho _{Sp}, \label{density}
\end{equation
are the baryon density and scalar density, respectively, with the latter
given by
\begin{eqnarray}
\rho _{SJ} &=&\frac{2}{(2\pi )^{3}}\int_{0}^{k_{F}^{J}}d\vec{k}\frac
M_{0}^{\ast }}{\sqrt{|\vec{k}|^{2}+{M_{0}^{\ast }}^{2}}} \notag \\
&=&\frac{M_{J}^{\ast }}{2\pi ^{2}}\left[ k_{F}^{J}E_{F}^{J\ast }-
M_{0}^{\ast 2}}\ln \frac{k_{F}^{J}+E_{F}^{J\ast }}{M_{0}^{\ast }}\right]
,J=p,n. \label{ScaDen}
\end{eqnarray
In the above expression, we have $E_{F}^{J\ast }=\sqrt{k_{F}^{J2}+M_{0}^
\ast 2}}$ with $M_{0}^{\ast }=M-g_{\sigma }\bar{\sigma}$ being the nucleon
Dirac mass and the Fermi momentum $k_{F}^{J}=k_{F}(1+\tau _{3}\delta )^{1/3}$
with $\tau _{3}=1$ for neutrons and $\tau _{3}=-1$ for protons, and
k_{F}=(3\pi ^{2}\rho /2)^{1/3}$ being the Fermi momentum for symmetric
nuclear matter.
The energy-momentum density tensor for the interacting Lagrangian density
\ref{NonRMF}) can be written as
\begin{equation}
\mathcal{T}^{\mu \nu }=\bar{\psi}i\gamma ^{\mu }\partial ^{\nu }\psi
+\partial ^{\mu }\sigma \partial ^{\nu }\sigma -F^{\mu \eta }\partial ^{\nu
}\omega _{\eta }-\vec{\mkern1muG}^{\mu \eta }\partial ^{\nu }\vec{\mker
1mu\rho }_{\eta }-\mathcal{L}g^{\mu \nu }, \label{EnMomTen}
\end{equation
where $g_{\mu \nu }=(+,-,-,-)$ is the Minkovski metric. In the mean field
approximation, the mean value of time (zero) component of the
energy-momentum density tensor is the energy density of the nuclear matter
system, i.e.
\begin{align}
\varepsilon =& \langle \mathcal{T}^{00}\rangle =\varepsilon _{\mathrm{kin
}^{n}+\varepsilon _{\mathrm{kin}}^{p}+\frac{1}{2}\left[ m_{\sigma }^{2}\bar
\sigma}^{2}+m_{\omega }^{2}\bar{\omega}_{0}^{2}+m_{\rho }^{2}\left( \bar{\rh
}_{0}^{(3)}\right) ^{2}\right] \notag \\
& +\frac{1}{3}b_{\sigma }(g_{\sigma }\bar{\sigma})^{3}+\frac{1}{4}c_{\sigma
}(g_{\sigma }\bar{\sigma})^{4}+\frac{3}{4}c_{\omega }(g_{\omega }\bar{\omega
_{0})^{4} \notag \\
& +\frac{1}{2}\left( g_{\rho }\bar{\rho}_{0}^{(3)}\right) ^{2}\left[ \Lambda
_{S}(g_{\sigma }\bar{\sigma})^{2}+3\Lambda _{V}(g_{\omega }\bar{\omega
_{0})^{2}\right] ,
\end{align
wher
\begin{eqnarray}
\varepsilon _{\mathrm{kin}}^{J} &=&\frac{2}{(2\pi )^{3}}\int_{0}^{k_{F}^{J}}
\vec{k}\sqrt{|\vec{k}|^{2}+{M_{J}^{\ast 2}}} \notag \\
&=&\frac{1}{\pi ^{2}}\int_{0}^{k_{F}^{J}}k^{2}dk\sqrt{k^{2}+{M_{0}^{\ast 2}}}
\notag \\
&=&\frac{1}{4}\left[ 3E_{F}^{J\ast }\rho _{J}+M_{0}^{\ast }\rho _{SJ}\right]
,~~J=p,n, \label{EnDenKin}
\end{eqnarray
is the kinetic part of the energy density. Similarly, the mean value of
space components of the energy-momentum density tensor corresponds to the
pressure of the system, i.e.
\begin{align}
P=& \frac{1}{3}\sum_{j=1}^{3}\langle \mathcal{T}^{jj}\rangle =P_{\mathrm{kin
}^{n}+P_{\mathrm{kin}}^{p} \notag \\
& -\frac{1}{2}\left[ m_{\sigma }^{2}\bar{\sigma}^{2}-m_{\omega }^{2}\bar
\omega}_{0}^{2}-m_{\rho }^{2}\left( \bar{\rho}_{0}^{(3)}\right) ^{2}\right]
\notag \\
& -\frac{1}{3}b_{\sigma }(g_{\sigma }\bar{\sigma})^{3}-\frac{1}{4}c_{\sigma
}(g_{\sigma }\bar{\sigma})^{4}+\frac{1}{4}c_{\omega }(g_{\omega }\bar{\omega
_{0})^{4} \notag \\
& +\frac{1}{2}\left( g_{\rho }\bar{\rho}_{0}^{(3)}\right) ^{2}\left[ \Lambda
_{S}(g_{\sigma }\bar{\sigma})^{2}+\Lambda _{V}(g_{\omega }\bar{\omega
_{0})^{2}\right] ,
\end{align
where the kinetic part of pressure is given by
\begin{equation}
P_{\mathrm{kin}}^{J}=\frac{1}{3\pi ^{2}}\int_{0}^{k_{F}^{J}}dk\frac{k^{4}}
\sqrt{k^{2}+{M_{0}^{\ast }}^{2}}},~~J=p,n. \label{pressureKin}
\end{equation}
The binding energy per nucleon of the asymmetric nuclear matter can be
calculated through the energy density by
\begin{equation}
E(\rho ,\delta )=\frac{\varepsilon (\rho ,\delta )}{\rho }-M. \label{SiEn}
\end{equation
Furthermore, the symmetry energy $E_{\mathrm{sym}}(\rho )$ can be obtained
as
\begin{equation}
E_{\mathrm{sym}}(\rho )=\left. \frac{1}{2!}\frac{\partial ^{2}E(\rho ,\delta
)}{\partial \delta ^{2}}\right\vert _{\delta =0}=\frac{k_{F}^{2}}
6E_{F}^{\ast }}+\frac{g_{\rho }^{2}\rho }{2Q_{\rho }}, \label{NonRMFEsym}
\end{equation
while the $4$th-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$ can be
expressed as
\begin{equation}
E_{\mathrm{{sym},4}}(\rho )=\left. \frac{1}{4!}\frac{\partial ^{4}E(\rho
,\delta )}{\partial \delta ^{4}}\right\vert _{\delta =0}=E_{\mathrm{{sym},4
}^{\mathrm{kin}}(\rho )+E_{\mathrm{{sym},4}}^{\mathrm{M}}(\rho ),
\label{NonRMFEsym4}
\end{equation
wher
\begin{align}
E_{\mathrm{{sym},4}}^{\mathrm{kin}}(\rho )=& \frac{k_{F}^{2}}{648}\frac{4
M_{0}^{\ast 4}}+11{M_{0}^{\ast 2}}k_{F}^{2}+10k_{F}^{4}}{E_{F}^{\ast 5}},
\label{Esym4Kin} \\
E_{\mathrm{{sym},4}}^{\mathrm{M}}(\rho )=& \frac{g_{\rho }^{8}\rho ^{3}}
2Q_{\rho }^{4}}\left( \frac{\Lambda _{V}^{2}g_{\omega }^{4}\bar{\omega
_{0}^{2}}{Q_{\omega }}-\frac{\Lambda _{S}^{2}g_{\sigma }^{4}\bar{\sigma}^{2
}{Q_{\sigma }}\right) \notag \\
& +\frac{g_{\sigma }^{2}\rho {M_{0}^{\ast }}k_{F}^{2}}{24Q_{\sigma
}E_{F}^{\ast 3}}\left( \frac{4\Lambda _{S}g_{\sigma }g_{\rho }^{4}\bar{\sigm
}\rho }{Q_{\rho }^{2}}-\frac{{M_{0}^{\ast }}k_{F}^{2}}{3E_{F}^{\ast 3}
\right) , \label{Esym4M}
\end{align
with $E_{\mathrm{{sym},4}}^{\mathrm{kin}}(\rho )$ representing the kinetic
part (including the interactions due to the nucleon effective mass) while
E_{\mathrm{{sym},4}}^{\mathrm{M}}(\rho )$ the other part due to the
interaction in the $4$th-order symmetry energy, and $E_{F}^{\ast }=\sqrt
M_{0}^{\ast 2}+k_{F}^{2}}$. The coefficients $Q_{\sigma }$, $Q_{\omega }$
and $Q_{\rho }$ are defined as
\begin{align}
Q_{\sigma }=& m_{\sigma }^{2}+g_{\sigma }^{2}\left( \frac{3\rho _{S}}{
M_{0}^{\ast }}}-\frac{3\rho }{E_{F}^{\ast }}\right) +2b_{\sigma }Mg_{\sigma
}^{3}\bar{\sigma}+3c_{\sigma }g_{\sigma }^{4}\bar{\sigma}^{2}, \label{Qf} \\
Q_{\omega }=& m_{\omega }^{2}+3c_{\omega }g_{\omega }^{4}\bar{\omega
_{0}^{2}, \label{Qw} \\
Q_{\rho }=& m_{\rho }^{2}+\Lambda _{S}g_{\sigma }^{2}g_{\rho }^{2}\bar{\sigm
}^{2}+\Lambda _{V}g_{\omega }^{2}g_{\rho }^{2}\bar{\omega}_{0}^{2}.
\label{Qr}
\end{align
In the above expressions, all the fields are calculated in the case of
symmetric nuclear matter, i.e., at $\delta =0$.
The analytical expression of the symmetry energy $E_{\mathrm{sym}}(\rho )$,
i.e., Eq. (\ref{NonRMFEsym}), is a well-known result firstly given in Ref.
\cite{Hor01}. To our best knowledge, the formulas (\ref{NonRMFEsym4})-(\re
{Qr}) give, for the first time, the analytical expression of the $4$th-order
symmetry energy $E_{\mathrm{{sym},4}}(\rho )$ in the RMF model, which are
the main results of the present work. These analytical expressions allow us
to evaluate accurately the $4$th-order symmetry energy $E_{\mathrm{{sym},4
}(\rho )$ and thus study the higher-order corrections to the empirical
parabolic approximation within the framework of the RMF model. Before
presenting numerical results, it is instructive to analyze firstly the low
density behavior of the $E_{\mathrm{{sym},4}}(\rho )$. When $\rho
\rightarrow 0$, the magnitude of all fields will approach to zero and both
E_{F}^{\ast }$ and $M_{0}^{\ast }$ will approach to $M$, leading to $E_
\mathrm{{sym},4}}^{\mathrm{M}}(\rho )\rightarrow 0$ from Eq. (\ref{Esym4M
) and $E_{\mathrm{{sym},4}}^{\mathrm{kin}}(\rho )\rightarrow \frac{1}{162
\frac{k_{F}^{2}}{M}$ from Eq. (\ref{Esym4Kin}). Therefore, in the
low density limit, we have
\begin{equation}
\lim_{\rho \rightarrow 0}E_{\mathrm{{sym},4}}(\rho )\rightarrow \frac{1}{162
\frac{k_{F}^{2}}{M}, \label{lowdenlimEsym4}
\end{equation
which is exactly the result from the free Fermi gas model as
expected.
\section{Results and Discussions}
\label{s3}
\subsection{The $4$th-order symmetry energy and higher-order effects on the
isobaric incompressibility of asymmetric nuclear matter}
\begin{figure}[tbh]
\includegraphics[scale=0.9]{Esym4rho.eps}
\caption{(Color online) Density dependence of the $4$th-order symmetry
energy $E_{\mathrm{{sym},4}}(\protect\rho )$ as well as its kinetic part $E_
\mathrm{{sym},4}}^{\mathrm{kin}}(\protect\rho )$ and interacting part $E_
\mathrm{{sym},4}}^{\mathrm{M}}(\protect\rho )$ from two accurately
calibrated interactions, i.e., FSUGold (a) and IU-FSU (b).}
\label{Esym4NL3FSUIUFSU}
\end{figure}
Shown in Fig. \ref{Esym4NL3FSUIUFSU} is the density dependence of the $4
th-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$ as well as its
kinetic part $E_{\mathrm{{sym},4}}^{\mathrm{kin}}(\rho )$ and interacting
part $E_{\mathrm{{sym},4}}^{\mathrm{M}}(\rho )$ using two accurately
calibrated interactions, i.e., FSUGold \cite{Tod05} and IU-FSU \cite{fatt10
. The FSUGold has been accurately calibrated to the ground-state properties
of closed-shell nuclei, their linear response, and the structure of neutron
stars while the IU-FSU is a recently developed effective interaction that
improves the FSUGold by incorporating some of the recent constraints on
properties of neutron stars. One can see from Fig. \ref{Esym4NL3FSUIUFSU}
that the $E_{\mathrm{{sym},4}}(\rho )$ is quite small (less than $0.7$ MeV)
at normal nuclear matter density while it increases with density and can
reach to about $7$ MeV at $\rho =1$ fm$^{-3}$. Furthermore, one can see that
the kinetic part $E_{\mathrm{{sym},4}}^{\mathrm{kin}}(\rho )$ dominates over
the interacting part $E_{\mathrm{{sym},4}}^{\mathrm{M}}(\rho )$ with the
latter is generally negative.
\begin{figure}[tbh]
\includegraphics[scale=0.9]{Esym42rho.eps}
\caption{(Color online) Ratio of the $4$th-order symmetry energy to the
symmetry energy as a function of density from two accurately calibrated
interactions, i.e., FSUGold and IU-FSU. }
\label{Esym4Esym2}
\end{figure}
In order to investigate higher-order $E_{\mathrm{{sym},4}}(\rho )$ effects
on the EOS of asymmetric nuclear matter, it may make more sense to calculate
the ratio of the $4$th-order symmetry energy to the symmetry energy, i.e.,
E_{\mathrm{{sym},4}}(\rho )/E_{\mathrm{sym}}(\rho )$. In Fig. \re
{Esym4Esym2}, we show this ratio as a function of density with interactions
FSUGold and IU-FSU. It is seen that the ratio $E_{\mathrm{{sym},4}}(\rho
)/E_{\mathrm{sym}}(\rho )$ has a very small value of about $2\%$ around
normal nuclear matter density $\rho _{0}$, but it can reach to about
6\%\sim 7\%$ at high densities (e.g., $1.0\,\mathrm{{fm}^{-3}}$ or $6\sim
7\rho _{0}$). This result is essentially consistent with the nonrelativistic
calculations in some phenomenological models \cite{XuC11}. These features
imply that the $E_{\mathrm{{sym},4}}(\rho )$ may become important at higher
densities in some extreme physical conditions such as in neutron star where
the isospin asymmetry $\delta $ can be close to unity. As an example, in the
next subsection, we shall study effects of the $4$th-order symmetry energy
on the proton fraction in $\beta $-stable neutron star matter. In addition,
one can see from Fig. \ref{Esym4Esym2} that in the low density limit, we
have $\lim\limits_{\rho \rightarrow 0}E_{\mathrm{{sym},4}}(\rho )/E_{\mathrm
sym}}(\rho )=1/27$ as expected from the free Fermi gas model.
\begin{table*}[tbh]
\caption{Characteristic parameters of asymmetric nuclear matter, namely,
\protect\rho _{0}$ ($\mathrm{{fm}^{-3}}$), $E_{0}(\protect\rho
_{0})$ (MeV),
$E_{\mathrm{sym}}(\protect\rho _{0})$ (MeV), $E_{\mathrm{{sym},4}}(\protec
\rho _{0})$ (MeV), $K_{0}$ (MeV), $J_{0}$ (MeV), $I_{0}$ (MeV), $L$ (MeV),
K_{\mathrm{sym}}$ (MeV), $J_{\mathrm{sym}}$ (MeV),
$I_{\mathrm{sym}}$ (MeV),
$L_{\mathrm{{sym},4}}$ (MeV), $K_{\mathrm{{sym},4}}$ (MeV), $J_{\mathrm{{sym
,4}}$ (MeV), $K_{\mathrm{asy}}$ (MeV), $K_{\mathrm{{sat},2}}$ (MeV), $K_
\mathrm{{sat},4}}$ (MeV), and the ratios $K_{\mathrm{{sat},2}}/K_{\mathrm{as
}}$ and $K_{\mathrm{{sat},4}}/K_{\mathrm{{sat},2}}$, for different interactions.
\begin{tabular}{|c|r|r|r|r|r|r|r|}
\hline
& FSUGold & IU-FSU & FSU-I & FSU-II & FSU-III & FSU-IV & FSU-V \\
\hline\hline
$\rho _{0}$ & $0.148$ & $0.155$ & $0.148$ & $0.148$ & $0.148$ & $0.148$ &
0.148$ \\ \hline $E_{0}(\rho _{0})$ & $-16.3$ & $-16.4$ & $-16.3$ &
$-16.3$ & $-16.3$ & $-16.3 $ & $-16.3$ \\ \hline
$E_{\mathrm{sym}}(\rho _{0})$ & $32.5$ & $31.3$ & $37.4$ & $35.5$ &
$33.9$ & $31.4$ & $30.9$ \\ \hline $E_{\mathrm{{sym},4}}(\rho _{0})$
& $0.66$ & $0.67$ & $0.66$ & $0.66$ & $0.66 $ & $0.66$ & $0.78$ \\
\hline $K_{0}$ & $229.2$ & $232.3$ & $229.2$ & $229.2$ & $229.2$ &
$229.2$ & $229.2$
\\ \hline
$J_{0}$ & $-521.6$ & $-288.5$ & $-521.6$ & $-521.6$ & $-521.6$ & $-521.6$ &
-521.6$ \\ \hline
$I_{0}$ & $2815.7$ & $4541.9$ & $2815.7$ & $2815.7$ & $2815.7$ & $2815.7$ &
2815.7$ \\ \hline $L$ & $60.4$ & $47.3$ & $109.5$ & $87.4$ & $71.7$
& $52.1$ & $49.4$ \\ \hline $K_{\mathrm{sym}}$ & $-51.4$ & $29.0$ &
$2.7$ & $-68.4$ & $-74.4$ & $-16.7$ & $5.5$ \\ \hline
$J_{\mathrm{sym}}$ & $426.5$ & $363.9$ & $-101.4$ & $157.6$ & $399.0$ &
251.2$ & $80.4$ \\ \hline $-I_{\mathrm{sym}}$ & $6331.8$ & $11346.5$
& $285.8$ & $1364.1$ & $4211.9$ & $6136.2$ & $4620.4$ \\ \hline
$L_{\mathrm{{sym},4}}$ & $1.9$ & $1.8$ & $1.9$ & $1.9$ & $1.9$ &
$1.9$ & $2.3 $ \\ \hline $K_{\mathrm{{sym},4}}$ & $0.5$ & $0.1$ &
$0.5$ & $0.5$ & $0.5$ & $0.5$ & $0.1 $ \\ \hline
$J_{\mathrm{{sym},4}}$ & $5.0$ & $6.3$ & $4.8$ & $4.8$ & $4.8$ &
$5.2$ & $5.1 $ \\ \hline
$K_{\mathrm{asy}}$ & $-413.8$ & $-255.1$ & $-654.3$ & $-592.6$ & $-504.4$ &
-329.4$ & $-290.9$ \\ \hline $K_{\mathrm{{sat},2}}$ & $-276.4$ &
$-196.3$ & $-405.1$ & $-393.8$ & $-341.4$ & $-210.8$ & $-178.5$ \\
\hline $K_{\mathrm{{sat},4}}$ & $3.0$ & $47.6$ & $338.6$ & $183.4$ &
$50.0$ & $12.9$ & $32.4$ \\ \hline
$K_{\mathrm{{sat},2}}/K_{\mathrm{asy}}$ & $67\%$ & $77\%$ & $62\%$ &
$66\%$ & $68\%$ & $64\%$ & $61\%$ \\ \hline
$-K_{\mathrm{{sat},4}}/K_{\mathrm{{sat},2}}$ & $1\%$ & $24\%$ & $84\%$ &
47\%$ & $15\%$ & $6\%$ & $18\%$ \\ \hline
\end{tabular
\label{TabKsat}
\end{table*}
The analytical expression of the $4$th-order symmetry energy $E_{\mathrm{{sy
},4}}(\rho )$ allows us to calculate accurately the density slope
and curvature parameters of $E_{\mathrm{{sym},4}}(\rho )$, i.e.,
$L_{\mathrm{{sym},4}}$ and $K_{\mathrm{{sym},4}}$, and thus obtain
the accurate value of the higher-order isobaric incompressibility
$K_{\mathrm{{sat},4}}$ of asymmetric nuclear matter according to Eq.
(\ref{Ksat4}). Table \ref{TabKsat} displays the characteristic
parameters of asymmetric nuclear matter, namely, $\rho
_{0}$, $E_{0}(\rho _{0})$, $E_{\mathrm{sym}}(\rho _{0})$, $E_{\mathrm{{sym},
}}(\rho _{0})$, $K_{0}$, $J_{0}$, $I_{0}$, $L$, $K_{\mathrm{sym}}$, $J_
\mathrm{sym}}$, $I_{\mathrm{sym}}$, $L_{\mathrm{{sym},4}}$, $K_{\mathrm{{sym
,4}}$, $J_{\mathrm{{sym},4}}$, $K_{\mathrm{asy}}$, $K_{\mathrm{{sat},2}}$,
K_{\mathrm{{sat},4}}$ and the ratios $K_{\mathrm{{sat},2}}/K_{\mathrm{asy}}$
and $K_{\mathrm{{sat},4}}/K_{\mathrm{{sat},2}}$, for the two accurately
calibrated interactions FSUGold and IU-FSU. To see the variation of the
higher-order characteristic parameters with the density dependence of the
symmetry energy, we also include in Table \ref{TabKsat} the results from $5$
interactions denoted as FSU-I, FSU-II, FSU-III, FSU-IV and FSU-V for which
the parameters $(\Lambda _{S},\Lambda _{V})$ are selected as $(0.00,0.00)$,
(0.00,0.01)$, $(0.00,0.02)$, $(0.00,0.04)$ and $(0.01,0.03)$, respectively,
while the $g_{\rho }$ parameter is adjusted accordingly to fix $E_{\mathrm
sym}}(\rho _{f})=25.57\, \mathrm{MeV}$ at $\rho _{f}=0.1\,\mathrm{{fm}^{-3}}$
as in the FSUGold interaction. The other parameters for FSU-I, FSU-II,
FSU-III, FSU-IV and FSU-V are exactly the same as in FSUGold (Note: the
FSUGold corresponds to the case of $(\Lambda _{S},\Lambda _{V})=(0.00,0.03)
). This is equivalent to solve a constraint equation about $g_{\rho
},\Lambda _{S}$ and $\Lambda _{V}$, i.e., $\frac{g_{\rho }^{2}\rho }
2Q_{\rho }}=13.47\,\mathrm{MeV}$ where $\bar{\sigma}$ and
$\bar{\omega}_{0}$ in $Q_{\rho }$ (See Eq. ((\ref{Qr}))) are
determined by the properties of symmetric nuclear matter in FSUGold.
From Table \ref{TabKsat}, one can see that the
$E_{\mathrm{{sym},4}}(\rho _{0})$ is generally less than $1$ MeV
(about $0.66$ MeV for most of the
interactions considered here), consistent with that observed in Fig. \re
{Esym4NL3FSUIUFSU}. These results about the $E_{\mathrm{{sym},4}}(\rho _{0})$
are further in agreement with the calculations in the nonrelativistic models
of MDI and Skyrme-Hartree-Fock \cite{Che09}, nicely verifying the empirical
parabolic law around the normal nuclear matter density $\rho _{0}$.
As pointed out previously, the difference between $K_{\mathrm{{sat},2}}$ and
$K_{\mathrm{asy}}$ reflects the contribution from higher-order effects,
namely, the value of ${J_{0}L}/{K_{0}}$, which has been usually neglected in
many calculations in the literature \cite{baran05,Che05,Cen09,LiT07,lopez88
. From Table \ref{TabKsat}, one can see that neglecting the ${J_{0}L}/{K_{0}}
$ term generally leads to $30$-$40\%$ relative error for the $K_{\mathrm{{sa
},2}}$ parameter and thus the ${J_{0}L}/{K_{0}}$ term contribution to the
K_{\mathrm{{sat},2}}$ parameter cannot be neglected simply, confirming the
previous findings in the nonrelativistic studies \cite{Che09}.
Furthermore, it is seen from Table \ref{TabKsat} that the value of
higher-order $K_{\mathrm{{sat},4}}$ is generally small compared with that of
$K_{\mathrm{{sat},2}}$ for most of the interactions considered here. In
addition, one can see that the $K_{\mathrm{{sat},4}}$ becomes more important
for the interactions with larger $L$ values and this is consistent with the
nonrelativistic studies \cite{Che09}. It should be noted that the
higher-order $K_{\mathrm{{sat},4}}$ term can be safely neglected in the
study of giant resonance of finite nuclei \cite{Pie09} where the isospin
asymmetry $\delta $ is usually small, i.e., about $0.2$.
\subsection{Effects of $E_{\mathrm{{sym},4}}$ on the proton fraction in
\protect\beta $-stable nuclear matter}
In order to further illustrate the effects of the $4$th-order symmetry
energy on the EOS of asymmetric nuclear matter, we calculate the proton
fraction $x_{p}$ in $\beta $-stable neutron star matter where the isospin
asymmetry $\delta $ is generally close to $1$. The chemical composition of
the neutron star is determined by the requirement of charge neutrality and
equilibrium with respect to the weak interaction ($\beta $-stable matter).
From the binding energy per nucleon, i.e., Eq. (\ref{EOSANM}), we can
calculate the proton fraction, $x_{p}=(1-\delta )/2$, for $\beta $-stable
nuclear matter as found in interior of neutron stars. For neutrino free
\beta $-stable nuclear matter, the chemical equilibrium for the reactions
n\rightarrow p+e^{-}+\bar{\nu}_{e}$ and $p+e^{-}\rightarrow n+\nu _{e}$
requires
\begin{eqnarray}
\mu _{e} &=&\mu _{n}-\mu _{p}=2\frac{\partial E}{\partial \delta } \notag \\
&=&4\delta E_{\mathrm{sym}}(\rho )+8\delta ^{3}E_{\mathrm{{sym},4}}(\rho )
\mathcal{O}(\delta ^{5}) \label{elechem1}
\end{eqnarray
where $\mu _{i}=\partial E_{i}/\partial x_{i}~(i=n,p,e,\mu )$ is the
chemical potential. For relativistic degenerate electrons, we have
\begin{eqnarray}
\mu _{e} &=&\left( m_{e}^{2}+k_{F}^{e2}\right) ^{1/2} \notag \\
&=&\left[ m_{e}^{2}+(3\pi ^{2}\rho x_{e})^{2/3}\right] ^{1/2}\simeq \left(
3\pi ^{2}\rho x_{e}\right) ^{1/3} \label{elechem2}
\end{eqnarray
where $m_{e}=0.511\,$MeV is the electron mass, and $x_{p}=x_{e}$ because of
charge neutrality.
Just above a nuclear matter density at which $\mu _{e}$ exceeds the muon
mass $m_{\mu }=0.105\,$GeV, the reactions $e^{-}\rightarrow \mu ^{-}+\nu
_{e}+\bar{\nu}_{\mu }$, $p+\mu ^{-}\rightarrow n+\nu _{\mu }$ and
n\rightarrow p+\mu ^{-}+\bar{\nu}_{\mu }$ are energetically allowed so that
both electrons and muons are present in $\beta $-stable nuclear matter, this
alters $\beta $-stability condition to
\begin{equation}
\mu _{n}-\mu _{p}=\mu _{e},~~\mu _{n}-\mu _{p}=\mu _{\mu }=\left[ m_{\mu
}^{2}+(3\pi ^{2}\rho x_{\mu })^{2/3}\right] ^{1/2} \label{elemuochem}
\end{equation
with $x_{p}=x_{e}+x_{\mu }$.
\begin{figure}[tbh]
\includegraphics[scale=0.86]{XpRho.eps}
\caption{(Color online) Density dependence of the proton fraction $x_{p}$ in
$\protect\beta $-stable $npe\protect\mu $ matter with FSUGold (a) and IU-FSU
(b). Three cases, i.e., the full EOS of asymmetric nuclear matter (solid
lines), its parabolic approximation (up to $\protect\delta ^{2}$ in Eq.
\protect\ref{EOSANM})) (dotted lines), and further including the $4$th-order
symmetry energy (up to $\protect\delta ^{4}$ in Eq. (\protect\ref{EOSANM}))
(dashed lines), are considered.}
\label{Xproton}
\end{figure}
In Fig. \ref{Xproton}, we show the density dependence of the proton fraction
$x_{p}$ in $\beta $-stable $npe\mu $ matter with the interactions FSUGold
and IU-FSU. We consider three cases for the EOS of asymmetric nuclear matter
here, i.e., the full one, its parabolic approximation (up to $\delta ^{2}$
in Eq. (\ref{EOSANM})), and\ the one further including the $4$th-order
symmetry energy (up to $\delta ^{4}$ in Eq. (\ref{EOSANM})). The results
show that, for both interactions of FSUGold and IU-FSU, the $4$th-order
symmetry energy is moderately important for the proton fraction, especially
at higher densities. For the FSUGold (IU-FSU)\ interaction and $\rho =1.0\
\mathrm{{fm}^{-3}}$, for instance, including the $4$th-order symmetry energy
in the parabolic approximation to the EOS of asymmetric nuclear matter will
increase the proton fraction $x_{p}$ from $15.37\%$ ($13.44\%$) to $16.43\%$
($14.81\%$), producing a relative variation of about $7\%$ ($10\%$). These
results indicate that the $4$th-order symmetry energy may have obvious
effects on the proton fraction $x_{p}$ in $\beta $-stable $npe\mu $ matter
and the parabolic approximation to the EOS of asymmetric nuclear matter may
significantly underestimate the proton fraction, especially at higher
densities. These features are consistent with the nonrelativistic
Skyrme-Hartree-Fock calculations \cite{Zha01}.
Furthermore, one can see from Fig. \ref{Xproton} that the difference between
the results with the full EOS and with the one containing the terms up to
the $4$th-order symmetry energy is very small, indicating that the EOS of
asymmetric nuclear matter including the terms up to the $4$th-order symmetry
energy (up to $\delta ^{4}$ in Eq. (\ref{EOSANM})) could be a good
approximation for the determination of the proton fraction in $\beta
-stable $npe\mu $ matter.
\subsection{Effects of $E_{\mathrm{{sym},4}}$ on core-crust transition
density and pressure in neutron stars}
The transition density $\rho _{t}$ is the baryon number density that
separates the liquid core from the inner crust in neutron stars and it plays
an important role in determining many properties of neutron stars \cit
{Hor01,Pro06,Duc08a,Duc08b,Lat07,xu09}. One simple and widely used way to
determine the core-crust transition density $\rho _{t}$ is the so-called
thermodynamical method, which requires the system to obey the following
intrinsic stability condition~\cite{Cal85,Kub07,Lat07}
\begin{eqnarray}
-\left( \frac{\partial P}{\partial v}\right) _{\mu _{np}} &>&0,
\label{ther1} \\
-\left( \frac{\partial \mu _{np}}{\partial q_{c}}\right) _{v} &>&0,
\label{ther2}
\end{eqnarray
where the $P=P_{b}+P_{e}$ is the total pressure of the $npe$ matter system
with $P_{b}$ and $P_{e}$ denoting the contributions from baryons and
electrons respectively, and the $v$ and $q_{c}$ are the volume and charge
per baryon number. The $\mu _{np}$ is defined as the chemical potential
difference between neutrons and protons, i.e., $\mu _{np}=\mu _{n}-\mu _{p}
. The pressure $P_{e}$ is only a function of the chemical potential
difference $\mu _{np}$ by assuming the $\beta $-equilibrium condition is
satisfied, i.e., $\mu _{np}=\mu _{e}$. By using the relation ${\partial
E_{b}(\rho ,x_{p})}/{\partial x_{p}}=-\mu _{np}$ with $E_{b}(\rho ,x_{p})$
being energy per baryon from the baryons in the $\beta $-equilibrium neutron
star matter and $x_{p}=\rho _{p}/\rho $, and treating the electrons as free
Fermi gas, one can show \cite{xu09} that the thermodynamical relations Eq.~
\ref{ther1}) and Eq.~(\ref{ther2}) are actually equivalent to the following
condition
\begin{widetext}
\begin{eqnarray}
V_{\mathrm{thermal}} &=&2\rho \frac{\partial E_{b}(\rho
,x_{p})}{\partial \rho }+\rho ^{2}\frac{\partial ^{2}E_{b}(\rho
,x_{p})}{\partial \rho ^{2}} -\left. \left( \frac{\partial
^{2}E_{b}(\rho ,x_{p})}{\partial \rho
\partial x_{p}}\rho \right) ^{2}\right/ \frac{\partial ^{2}E_{b}(\rho ,x_{p}
}{\partial x_{p}^{2}}>0, \label{Vther}
\end{eqnarray
which determines the thermodynamical instability region of the $\beta
-equilibrium neutron star matter. The baryon number density that violates
the condition Eq.~(\ref{Vther}) then corresponds to the core-crust
transition density in neutron stars for the thermodynamical method.
With the EOS of asymmetric nuclear matter including the terms up to the $4
th-order symmetry energy (i.e., up to $\delta ^{4}$ in Eq. (\ref{EOSANM})),
Eq.~(\ref{Vther}) is then reduced to
\begin{eqnarray}
V_{\mathrm{thermal}} &=&\rho ^{2}\frac{\partial ^{2}E_{0}}{\partial \rho ^{2
}+2\rho \frac{\partial E_{0}}{\partial \rho }+\delta ^{2}\left[ \rho ^{2
\frac{\partial ^{2}E_{\mathrm{sym}}(\rho )}{\partial \rho ^{2}}+2\rho \frac
\partial E_{\mathrm{sym}}(\rho )}{\partial \rho }\right] +\delta ^{4}\left[
\rho ^{2}\frac{\partial ^{2}E_{\mathrm{sym,4}}(\rho )}{\partial \rho ^{2}
+2\rho \frac{\partial E_{\mathrm{sym,4}}(\rho )}{\partial \rho }\right]
\notag \\
&&-\frac{\rho ^{2}\delta ^{2}}{E_{\mathrm{sym}}(\rho )+6E_{\mathrm{sym,4
}(\rho )}\left[ \frac{\partial E_{\mathrm{sym}}(\rho )}{\partial \rho
+2\delta ^{2}\frac{\partial E_{\mathrm{sym,4}}(\rho )}{\partial \rho }\right]
^{2}>0. \label{Vther4th}
\end{eqnarray
The baryon number density that violates the condition Eq.~(\re
{Vther4th}) then corresponds to the core-crust transition density
$\rho _{t}^{\mathrm{4th}}$ in neutron stars for the EOS of
asymmetric nuclear
matter including the terms up to the $4$th-order symmetry energy (up to
\delta ^{4}$ in Eq. (\ref{EOSANM})). The corresponding transition pressure
P_{t}^{\mathrm{4th}}$ at $\rho _{t}^{\mathrm{4th}}$ for the EOS of
asymmetric nuclear matter including the terms up to the $4$th-order
symmetry energy (up
to $\delta ^{4}$ in Eq. (\ref{EOSANM})) is then given b
\begin{eqnarray}
P_{t}^{\mathrm{4th}}
&=&P_{t}^{b,\mathrm{4th}}+P_{t}^{e,\mathrm{4th}} \notag
\\
&=&\left[ \rho ^{2}\left( \frac{\partial E_{0}}{\partial \rho }+\delta ^{2
\frac{\partial E_{\mathrm{sym}}}{\partial \rho }+\delta ^{4}\frac{\partial
E_{\mathrm{sym,4}}}{\partial \rho }\right) \right] _{\rho =\rho _{t}}+\mu
_{e}^{\mathrm{4th}}\rho _{e} \notag \\
&=&\left[ \rho ^{2}\left( \frac{\partial E_{0}}{\partial \rho }+\delta ^{2
\frac{\partial E_{\mathrm{sym}}}{\partial \rho }+\delta ^{4}\frac{\partial
E_{\mathrm{sym,4}}}{\partial \rho }\right) \right] _{\rho =\rho
_{t}}+\left\{ 4\rho \delta \cdot \frac{1-\delta }{2}\cdot \lbrack E_{\mathrm
sym}}+\delta ^{2}E_{\mathrm{sym,4}}]\right\} _{\rho =\rho _{t},\delta
=\delta _{t}}
\end{eqnarray
where $\delta _{t}$ is the isospin asymmetry of the $\beta
-equilibrium neutron star matter at the corresponding transition density.
The transition density obtained by neglecting the $E_{\mathrm{sym,4}}(\rho )$
term in the condition Eq.~(\ref{Vther4th}) corresponds to the core-crust
transition density $\rho _{t}^{\mathrm{2nd}}$ in neutron stars for the
parabolic approximation to the EOS of asymmetric nuclear matter (up to
\delta ^{2}$ in Eq. (\ref{EOSANM})). The corresponding transition
pressure at $P_{t}^{\mathrm{2nd}}$ for the parabolic approximation
to the EOS of asymmetric nuclear matter (up to $\delta ^{2}$ in Eq.
(\ref{EOSANM})) is then
expressed a
\begin{eqnarray}
P_{t}^{\mathrm{2nd}}
&=&P_{t}^{b,\mathrm{2nd}}+P_{t}^{e,\mathrm{2nd}} \notag
\\
&=&\left[ \rho ^{2}\left( \frac{\partial E_{0}}{\partial \rho }+\delta ^{2
\frac{\partial E_{\mathrm{sym}}}{\partial \rho }\right) \right] _{\rho =\rho
_{t}}+\mu _{e}^{\mathrm{2nd}}\rho _{e}=\left[ \rho ^{2}\left( \frac{\partial
E_{0}}{\partial \rho }+\delta ^{2}\frac{\partial E_{\mathrm{sym}}}{\partial
\rho }\right) \right] _{\rho =\rho _{t}}+\left[ 4\rho \delta \cdot \frac
1-\delta }{2}\cdot E_{\mathrm{sym}}\right] _{\rho =\rho _{t},\delta =\delta
_{t}}.
\end{eqnarray
\end{widetext}
Due to simplicity, the $\rho _{t}^{\mathrm{2nd}}$ and $P_{t}^{\mathrm{2nd}}$
have been extensively applied to determine the inner edge of neutron star
crusts within the nonrelativistic models \cite{Kub07,Lat07,Oya07,Wor08} and
recently in the RMF model \cite{Mou10} as well. However, recent studies
based on some nonrelativistic models have demonstrated \cite{xu09} that the
parabolic approximation to the EOS of asymmetric nuclear matter may lead
systematically to significantly higher core-crust transition densities and
pressures, especially with stiffer symmetry energy functionals. It is thus
very interesting to see how the higher-order $E_{\mathrm{sym,4}}(\rho )$
affects the transition density $\rho _{t}$ and pressure $P_{t}$ in the RMF
model.
In Table \ref{rhoT}, we show the $\rho _{t}^{\mathrm{2nd}}$, $\rho _{t}^
\mathrm{4th}}$, $P_{t}^{\mathrm{2nd}}$, and $P_{t}^{\mathrm{4th}}$ obtained
from the thermodynamical method with different interactions as in In Table
\ref{TabKsat}. It is interesting to see that including the $4$th-order
symmetry energy in the parabolic approximation to the EOS of asymmetric
nuclear matter indeed reduces significantly the core-crust transition
density $\rho _{t}$, which is consistent with the nonrelativistic
calculations \cite{xu09}. Furthermore, one can see that the $4$th-order
symmetry energy may have even more drastic effects on the core-crust
transition pressure $P_{t}$, namely, including the $4$th-order symmetry
energy in the parabolic approximation to the EOS of asymmetric nuclear
matter reduces drastically the core-crust transition pressure $P_{t}$.
Therefore, our results indicate that the empirical parabolic approximation
may cause large errors for the determination of the $\rho _{t}$ and $P_{t}$
in neutron stars in the nonlinear RMF model.
\begin{table*}[tbh]
\caption{The $\protect\rho _{t}^{\mathrm{2nd}}$ (fm$^{-3}$), $\protect\rho
_{t}^{\mathrm{4th}}$ (fm$^{-3}$), $P_{t}^{\mathrm{2nd}}$ (MeV/fm$^{3}$), and
$P_{t}^{\mathrm{4th}}$ (MeV/fm$^{3}$) obtained from the thermodynamical
method with different interactions.}
\label{rhoT}\centerin
\begin{tabular}{|c|r|r|r|r|r|r|r|}
\hline
& FSUGold & IU-FSU & FSU-I & FSU-II & FSU-III & FSU-IV & FSU-V \\
\hline\hline
$\rho _{t}^{\mathrm{2nd}}$ & $0.089$ & $0.090$ & $0.085$ & $0.088$ & $0.088$
& 0.083 & 0.080 \\ \hline
$\rho _{t}^{\mathrm{4th}}$ & $0.051$ & $0.077$ & $0.069$ & $0.054$ & $0.053$
& $0.068$ & 0.072 \\ \hline
$P_{t}^{\mathrm{2nd}}$ & $1.316$ & $0.673$ & $0.664$ & $1.010$ & $0.968$ &
0.621 & 0.501 \\ \hline
$P_{t}^{\mathrm{4th}}$ & $0.321$ & $0.530$ & $0.302$ & $0.236$ & $0.259$ &
0.420 & 0.414 \\ \hline
\end{tabular
\end{table*}
\section{Summary}
\label{s4}
We have derived for the first time the analytical expression of the
nuclear matter fourth-order symmetry energy
$E_{\mathrm{{sym},4}}(\rho )$ within the framework of the nonlinear
RMF model. It should be mentioned that the analytical expression of
$E_{\mathrm{{sym},4}}(\rho )$ can be easily generalized to the case
of the density dependent RMF model that has similar isospin
structure as the nonlinear RMF model (See, e.g., Ref.~\cite{Che07}).
This provides the possibility to investigate the higher-order
$E_{\mathrm{{sym},4}}(\rho )$ corrections to the widely used
empirical parabolic law for the asymmetric nuclear matter in the RMF
model. In the present work, as examples, we have investigated the
$E_{\mathrm{{sym},4}}(\rho )$ effects on the properties of
asymmetric nuclear matter, the proton fraction $x_{p}$ in $\beta $-stable
npe\mu $ matter and the core-crust transition density $\rho _{t}$ and
pressure $P_{t}$ in neutron stars within the nonlinear RMF model with two
accurately calibrated interactions, i.e., FSUGold and IU-FSU.
Firstly, our results have indicated that the value of $E_{\mathrm{{sym},4
}(\rho )$ at normal nuclear matter density $\rho _{0}$ is generally less
than $1$ MeV, and thus the empirical parabolic approximation $E(\rho ,\delta
)\simeq E_{0}(\rho )+E_{\mathrm{sym}}(\rho )\delta ^{2}$ has been nicely
confirmed around $\rho _{0}$. However, at higher densities such as $1$ fm
^{-3}$, the value of $E_{\mathrm{{sym},4}}(\rho )$ can be about $7$ MeV and
the ratio of $E_{\mathrm{{sym},4}}(\rho )/E_{\mathrm{sym}}(\rho )$ can reach
to about $7\%$. These results imply that the $E_{\mathrm{{sym},4}}(\rho )$
may become nonnegligible at higher densities. Furthermore, the analytical
form of the $E_{\mathrm{{sym},4}}(\rho )$ allows us to study the
higher-order effects on the isobaric incompressibility of asymmetric nuclear
matter. Our results have indicated that the value of higher-order $K_
\mathrm{{sat},4}}$ is generally small compared with that of $K_{\mathrm{{sat
,2}}$, confirming the previous nonrelativistic calculations \cite{Che09}.
Secondly, for the proton fraction $x_{p}$ in $\beta $-stable $npe\mu $
matter, we have found that, compared with the results from the empirical
parabolic approximation to the EOS of asymmetric nuclear matter, including
the $4$th-order symmetry energy $E_{\mathrm{{sym},4}}(\rho )$ can enhance
the proton fraction $x_{p}$ by about $10\%$ at higher densities. These
results indicate that the empirical parabolic approximation to the EOS of
asymmetric nuclear matter may cause obvious errors for the determination of
the proton fraction in neutron stars within the nonlinear RMF model, which
is in agreement with the results from the nonrelativistic models \cite{Zha01
.
Finally, we have demonstrated that including the $4$th-order symmetry energy
$E_{\mathrm{{sym},4}}(\rho )$ in the parabolic approximation to the EOS of
asymmetric nuclear matter can reduce significantly the core-crust transition
density $\rho _{t}$ and furthermore it has even more drastic effects on the
core-crust transition pressure $P_{t}$. Therefore, our results have clearly
demonstrated that the extensively used empirical parabolic approximation to
the EOS of asymmetric nuclear matter may lead systematically to
significantly higher core-crust transition density $\rho _{t}$ and pressure
P_{t}$ in neutron stars within the nonlinear relativistic mean field model,
confirming the previous finding based on nonrelativistic calculations \cit
{xu09}.
Therefore, we conclude that the higher-order
$E_{\mathrm{{sym},4}}(\rho )$ in the EOS of asymmetric nuclear
matter may have different effects on different quantities, and
generally one cannot simply neglect them, especially under some
extreme physical conditions, such as in neutron stars.
\section*{ACKNOWLEDGMENTS}
This work was supported in part by the National Natural Science
Foundation of China under Grant Nos. 10975097 and 11135011, the
Shanghai Rising-Star Program under grant No. 11QH1401100, ``Shu
Guang" project supported by Shanghai Municipal Education Commission
and Shanghai Education Development Foundation, the Program for
Professor of Special Appointment (Eastern Scholar) at Shanghai
Institutions of Higher Learning, and the National Basic Research
Program of China (973 Program) under Contract No. 2007CB815004.
|
\section{Introduction}
A polynomial matrix is a matrix whose entries belong to some polynomial ring $R$ \cite{tom}. In this paper we will always assume that
$R$ is a principal ideal domain. This condition is equivalent to $R=\mathbb{F}[x]$, the ring of univariate polynomials in $x$ with coefficients lying in some field $\mathbb{F}$.
An important property of a polynomial matrix with entries in $\mathbb{F}[x]$ is its complete eigenstructure, whose definition is given in Subsection \ref{completeigenstructure}. The name \emph{eigenstructure} comes from the special case where $\mathbb{F}=\mathbb{C}$; in this context, polynomial matrices are usually seen instead as matrix polynomials, that is polynomials whose coefficients are matrices \cite{bible}. Any matrix polynomial is associated with a polynomial eigenvalue problem (PEP); the complete eigenstructure is strictly related with the properties of the associated PEP. More precisely, it gives the complete information about the eigenvalues, eigenvectors and Jordan chains of the matrix polynomial, and also about the Kronecker form of any strong linearization of the matrix polynomial \cite{dopico}. Polynomial eigenvalue problems arise in many applications, from mathematics, science and engineering; both their algebraic properties and the numerical methods for their approximate solutions are widely studied. See, e.g., \cite{golub, tisseur, mehrmannvoss}.
The aim of this paper is to investigate the link between the complete eigenstructures of two polynomial matrices $P(x)$ and $Q(y)$
related one to another by a rational transformation $x(y)$ of the variable. In order to better explain the question we are interested
in, let us consider the following example, where $R=\mathbb{C}[x]$. Suppose that we have to deal with the polynomial matrix \[P(x)=
\left[\begin{array}{ccc} x^2-20x & 0 & 0\\x-20 &
x^2-20x & 0\\0 & 0 & x\\0 & 0 & x^2\\0 & 0 & 0\end{array}\right];
\]
if we choose grade($P(x)$)$=2$ (the grade of $P(x)$ is an arbitrary integer $g$ such that $g \geq \deg P(x)$; more details are given in Section \ref{definiz}), then the complete eigenstructure of $P(x)$ is the following:
\begin{itemize}
\item the elementary divisors of $P(x)$ are $(x-20)$, $x$, $(x-20)$, $x^2$;
\item there are no right minimal indices;
\item the left minimal indices of $P(x)$ are $0, \ 1$.
\end{itemize}
The rational change of variable $x(y)=\frac{16 y^2 - 25}{y^2-y}$ induces an application $\Phi_2$, as defined in \eqref{rationalfunction}, such that $\Phi_2(P(x))=(y^2-y)^2 P(\frac{16 y^2 - 25}{y^2-y})=:Q(y)$, with grade($Q(y))=4$ (see Section \ref{rationalfun}) and
\[Q(y)=
\left[\begin{array}{cccc} (25-16y^2)(2y-5)^2 & 0 & 0\\(y-y^2)(2y-5)^2 &
(25-16y^2)(2y-5)^2 & 0\\0 & 0 & (y^2-y)(16y^2-25)\\0 & 0 & (16y^2-25)^2\\0 & 0 & 0\end{array}\right].
\]
By studying the complete eigenstructure of $Q(y)$ we find out that
\begin{itemize}
\item the elementary divisors of $Q(y)$ are $(y-\frac{5}{2})^2$, $(y-\frac{5}{4})$, $(y+\frac{5}{4})$, $(y-\frac{5}{2})^2$, $(y-\frac{5}{4})^2$, $(y+\frac{5}{4})^2$;
\item there are no right minimal indices;
\item the left minimal indices of $Q(y)$ are $0, \ 2$.
\end{itemize}
Notice that $x(\frac{5}{2})=20$, $x(\pm \frac{5}{4})=0$, and that $y=\frac{5}{2}$ is a root of multiplicity $2$ of the equation
$x(y)=20$ while $y=\pm \frac{5}{4}$ are roots of multiplicity $1$ of the equation $x(y)=0$. We can therefore conjecture that if
$(x-x_0)^{\ell}$ is an elementary divisor of $P(x)$ and $y_0$ is a root of multiplicity $m$ of the equation
$x(y)=x_0$ then $(y-y_0)^{m \cdot \ell}$ is an elementary divisor of $Q(y)$. Moreover, we see that apparently the minimal indices have
been multiplied by a factor $2$; notice that $2$ is the degree of the considered rational transformation (that is the maximum of the
degrees of the numerator and the denominator).
The main result of the present paper is the proof that the conjectures above, which will be stated more precisely in Section \ref{themainresult}, are true for every rational transformation of the variable $x(y)$ and every polynomial matrix $P(x)$. Moreover, analogous properties hold for infinite elementary divisors and right minimal indices.
The motivation for this work comes from the will to generalise the partial results derived in \cite{dicksonpal}, where we considered the particular case of a square and regular polynomial matrix with entries in $\mathbb{C}[x]$ and without infinite elementary divisors, and the Dickson change of variable $x(y)=\frac{y^2+1}{y}$. Moreover, we wish to extend the results by D. S. Mackey and N. Mackey \cite{iciam}, who described the special case of rational transformations of degree $1$, also known as M\"{o}bius transformations. The present contribution is offered as both a synthesis and an extension of the previous works cited above.
The results provided in this paper can be used to design numerical methods for the approximate solution of PEPs. An example in this regard, restricted to the case of the Dickson transformation, is given in \cite{dicksonpal} for the solution of the palindromic PEP.
The structure of this paper is the following: in Section \ref{definiz} we expose the theoretical background we are going to work within,
and we give some basic definitions that we will use later on. In Section \ref{rationalfun} we formally define the application between
polynomial matrices induced by a rational change of variable and we present some intermediate results. Our main result is Theorem
\ref{main}, which is stated and commented in Section \ref{themainresult}; Sections \ref{proof1} and \ref{proof2} are devoted to the
proof of our result.
For the sake of simplicity, in Sections \ref{proof1} and \ref{proof2} we assume that the underlying field is
algebraically closed: in Section \ref{comments} we show how the result still holds for an arbitrary field.
Finally, in Section
\ref{app}, root polynomials are introduced in order to prove a technical Lemma.
The first part of Theorem
\ref{main} was stated and proved, but only for a very special case, in \cite{dicksonpal}. Besides the generalisation to a generic
rational transformation and a generic polynomial matrix, this paper also contains the analysis of what happens to minimal indices
and infinite elementary divisors.
\section{Preliminary definitions}\label{definiz}
In this Section we describe our notation and recall some basic definitions.
\subsection{Basic facts on polynomials}\label{gradi}
Let $Z$ be a ring and let $Z[x]$ be the ring of the univariate polynomials in the variable $x$ with coefficients in $Z$.
We denote the degree of $z \in Z[x]$ by the letter $k$, and sometimes
write $k=\deg z$.
On the other hand, the \emph{grade} \cite{m4} of a polynomial $z \in Z[x]$ is any integer $g=\mathrm{grade}(z)$ satisfying $g \geq k$.
The choice of the grade of a polynomial is arbitrary: nevertheless, some algebraic properties of polynomial matrices depend on the grade.
\begin{remark}
In some sense, the degree of a polynomial is an intrinsic property while the grade depends on its representation.
In fact, informally speaking, the grade depends on how many zero coefficients one wishes to add in front of the polynomial.
\end{remark}
Let now $g$ be the grade of $z=\sum_{i=0}^g a_i x^i \in Z[x]$. The \emph{reversal} of $z$ with respect to its grade \cite{gkl, m4} is
\begin{equation}\label{rev}
\mathrm{Rev}_g z := \sum_{i=0}^g a_{g-i} x^i.
\end{equation}
The subscript $g$ will sometimes be omitted when the reversal is taken with respect to the degree of the polynomial, that is $\mathrm{Rev}_k z=:\mathrm{Rev} z$.
Let now $\mathbb{F}$ be an arbitrary algebraically closed field.
\begin{remark}
Although the hypothesis that $\mathbb{F}$ is algebraically closed is useful to state in a simpler way our results, it is not strictly necessary. See Section \ref{comments}.
\end{remark}
A well-known result that is crucial to us is that $\mathbb{F}[x]$ is guaranteed to be an Euclidean domain.
Given $z_1, z_2 \in \mathbb{F}[x]$, not both zero, we denote by $\mathrm{GCD}(z_1,z_2)$ their greatest common divisor; we additionally require that $\mathrm{GCD}(z_1,z_2)$ is always monic so that it is uniquely defined.
We say that $z_1$ and $z_2$ are \emph{coprime} if $\mathrm{GCD}(z_1,z_2)=1_{\mathbb{F}[x]}$.
Notice that a polynomial $z \in \mathbb{F}[x]$ can be thought of as a function
$z(x):\mathbb{F}\rightarrow\mathbb{F}$. Thus, applying \eqref{rev}, in this case the formula $\mathrm{Rev}_gz(x)=x^g z(x^{-1})$ holds.
Let now $Z^{m \times p}$ be the set of $m \times p$
matrices with entries in $Z$; the case $p=1$ corresponds to the set of vector with
$m$ elements in $Z$, denoted by $Z^m$. We are mainly interested in analysing $(\mathbb{F}[x])^{m \times p}$,
the set of $m \times p$ \emph{polynomial matrices}
with entries in $\mathbb{F}[x]$.
$M_m(\mathbb{F}[x]):=(\mathbb{F}[x])^{m \times m}$ is the \emph{ring of square polynomial matrices of dimension $m$}.
A square polynomial matrix $A \in M_m(\mathbb{F}[x])$ is said to be \emph{regular} if $\det A \neq 0_{\mathbb{F}[x]}$ and \emph{singular} otherwise. If $A$ is regular and $\det A \in \mathbb{F}$ then $A$ is called \emph{unimodular}.
\begin{remark}
Notice that $(\mathbb{F}[x])^{m \times p}=(\mathbb{F}^{m \times p})[x]$; or in other words, a polynomial matrix, defined as a matrix whose entries are polynomials, is also a matrix polynomial, defined as a polynomial whose coefficients are matrices.
\end{remark}
The notions of grade and degree can be extended in a straightforward way to polynomial matrices, as follows: the grade (resp., the degree) of $A \in (Z[x])^{m \times p}$ is defined as $\max_{i,j}$grade$(A_{ij})$ (resp., as $\max_{i,j} \deg A_{ij}$). Analogously, the reversal of a polynomial matrix is defined just as in \eqref{rev}, after replacing $a_i \in Z$ with $B_i \in Z^{m \times p}$.
\subsection{Characteristic values, elementary divisors, and minimal indices}\label{completeigenstructure}
Let $A \in (F[x])^{m \times p}$, and let $\nu=:\min(m,p)$. Suppose that there exist $D_1, \dots, D_{\nu} \in \mathbb{F}[x]$ such that
$A_{ij}=D_i \delta_{ij}$, where $\delta_{ij}$ is the Kronecker's delta.
Then we write $A=\mathrm{diag}(D_{1},\ \dots, \ D_{\nu})$, and we say that $A$ is \emph{diagonal}. Notice that we use the notation
indifferently for both square and rectangular polynomial matrices.
The following Theorem, which in its most general is due to Frobenius \cite{frobenius}, is in point of fact valid for any
matrix with entries in a principal ideal domain \cite{tom, bible}.
\begin{theorem}\label{smiththeorem}
Let $P(x) \in (\mathbb{F}[x])^{m \times p}$.
Then there exist two unimodular $A(x) \in M_m(\mathbb{F}[x])$ and $B(x) \in M_p(\mathbb{F}[x])$ such that
\[S(x)=A(x)P(x)B(x)=\mathrm{diag}(d_1(x),\dots,d_{\nu}(x)),\] where $d_i(x) \in \mathbb{F}[x]$ is monic
$\forall i \leq \nu:=\min(m,p)$ and
$d_{i}(x)|d_{i+1}(x) \ \ \forall i \leq \nu-1$.
\end{theorem}
Such an $S(x) \in (\mathbb{F}[x])^{m \times p}$ is called the \emph{Smith form} \cite{bible, smith} of $P(x)$,
and the $d_i(x)$ are called its \emph{invariant polynomials} \cite{tom, bible}. The Smith form, and thus the invariant
polynomials, are uniquely determined by $P(x)$. Notice that a square polynomial matrix $P(x)$ is singular if and only if at least one
of its invariant polynomials is zero.
Using the fact that $\mathbb{F}$ is algebraically closed, let us consider a factorization of the invariant polynomials over
$\mathbb{F}[x]$: $d_i(x)=\prod_j (x-x_j)^{k_{j,(i)}}$.
The factors $(x-x_j)^{k_{j,(i)}}$ are called the \emph{elementary divisors} of $P(x)$ \cite{tom, bible} corresponding to the
\emph{characteristic value} $x_j$ \cite{tom}. Notice that, from Theorem \ref{smiththeorem},
$i_1\leq i_2 \Rightarrow k_{j,(i_1)}\leq k_{j,(i_2)}$.
\begin{remark}
When $\mathbb{F}=\mathbb{C}$ the characteristic values of the polynomial matrix $P(x)$ are often
called the eigenvalues of the matrix polynomial $P(x)$. Given an eigenvalue $x_0$, there is a Jordan chain of length $\ell$ at
$x_0$ if and only if $(x-x_0)^{\ell}$ is an elementary divisor. The number of Jordan chains at $x_0$ is equal to the number of invariant
polynomials that have $x_0$ as a root \cite{bible}.
\end{remark}
Let us now denote by $\mathbb{F}(x)$ the \emph{field of fractions} of the ring $\mathbb{F}[x]$.
Let $\mathcal{V}$ be a vector subspace of $(\mathbb{F}(x))^p$, with $\dim \mathcal{V}=s$. Let $\{v_i\}$ be a polynomial basis
for $\mathcal{V}$ with the property $\deg v_1 \leq \dots \leq \deg v_s$. Often we will arrange a polynomial basis in the matrix form
$V(x)=[v_1(x),\dots,v_s(x)] \in (\mathbb{F}[x])^{p \times s}$. Clearly, polynomial bases always exist, because one may start from any
basis with elements in the (vectorial) field of fractions, and then build a polynomial basis just by multiplying by the least common denominator. Let $\alpha_i:=\deg v_i$ be the degrees of the vectors of such a polynomial basis; the \emph{order} of $V(x)$ is defined \cite{forneyjr} as $\sum_{i=1}^s \alpha_i$. A polynomial basis is called \emph{minimal} \cite{forneyjr} if its order is minimal amongst all the polynomial bases for $\mathcal{V}$, and the $\alpha_i$ are called its \emph{minimal indices} \cite{forneyjr}. It is possible to prove \cite{forneyjr, tom} that, although there is not a unique minimal basis, the minimal indices are uniquely determined by $\mathcal{V}$.
The \emph{right minimal indices} \cite{dopico} of a polynomial matrix $P(x) \in (\mathbb{F}[x])^{m \times p}$ are defined as the minimal indices of $\ker P(x)$. Analogously, the \emph{left minimal indices} \cite{dopico} of $P(x)$ are the minimal indices of $\ker P(x)^T$.
Given the grade $g$ of $P(x)$, we say that $\infty$ is a characteristic value of $P(x)$ if $0_{\mathbb{F}}$ is a characteristic value of $Rev_g P(x)$. The elementary divisors corresponding to $\infty$ are defined \cite{diciotto} as the elementary divisors of $Rev_gP(x)$ corresponding to $0_{\mathbb{F}}$; if $x^{\ell}$ is an elementary divisor of $Rev_gP(x)$ we formally write that $(x-\infty)^{\ell}$ is an infinite elementary divisor of $P(x)$. Notice that the infinite elementary divisors of a polynomial matrix clearly depend on the arbitrary choice of its grade.
We complete this section with the following definition \cite{dopico}: the \emph{complete eigenstructure} of $P(x)$ is the set of both finite and infinite elementary divisors of $P(x)$ and of its left and right minimal indices.
\section{Rational transformations of polynomial matrices}\label{rationalfun}
Let $n(y), d(y) \in \mathbb{F}[y]$ be two nonzero, coprime polynomials. Let us define $N := \deg n(y)$, $D := \deg d(y)$,
and $G:=\max(N,D)$. We will always suppose $G \geq 1$, that is $n(y)$ and $d(y)$ are not both elements of $\mathbb{F}$. We denote
the coefficients of $n(y)$ and $d(y)$ as $n_i \in \mathbb{F}$, $i=0,\dots,N$ and $d_j \in \mathbb{F}$, $j=0,\dots,D$, that is
$n(y)=\sum_{i=0}^N n_i y^i$, $d(y)=\sum_{i=0}^D d_i y^i$.
Let us introduce the notation $\mathbb{F}^*:=\mathbb{F} \cup \{\infty\}$, having formally defined $\infty:=0_{\mathbb{F}}^{-1}$.
We consider the generic rational function from $\mathbb{F}^*$ to $\mathbb{F}^*$:
\begin{equation}\label{rationalfunction}
x(y)=\frac{n(y)}{d(y)}.
\end{equation}
The function \eqref{rationalfunction} induces an application $\Phi_{g,n(y),d(y)}:(\mathbb{F}[x])^{m \times p} \rightarrow (\mathbb{F}[y])^{m \times p}$ defined as
\begin{equation}\label{transformation}
\Phi_{g,n(y),d(y)}(P(x)) = Q(y):=[d(y)]^g P(x(y))
\end{equation}
Here $g$ is the grade of $P(x) \in (\mathbb{F}[x])^{m \times p}$, so for any choice of $g$ a different application is defined. We will usually omit the functional dependence of $\Phi$ on $n(y)$ and $d(y)$ unless the context allows any possible ambiguity; also, if the grade is chosen to be $g=k$ we will sometimes omit the subscript $g$, that is $\Phi(P(x)):=\Phi_{k,n(y),d(y)}(P(x))$.
Since a polynomial matrix is also a matrix polynomial, we can write $P(x)=\sum_{i=0}^g P_i x^i$ for some $P_i \in \mathbb{F}^{m \times p}$, $i=0,\dots,g$. Notice that following the same point of view we can also write $Q(y)=\sum_{i=0}^{g} P_i [n(y)]^i [d(y)]^{g-i}$.
\begin{lemma}\label{degQ}
$\deg Q(y)=\deg \Phi_g(P(x))$ is less than or equal to $q:=$
$gD+\max_{i : P_i \neq 0}(iN-iD)$. If $N \neq D$ the strict equality $\deg Q(y)=q$ always holds. Moreover, $q \leq gG$.
\end{lemma}
\begin{proof}
Writing $Q(y)$ as above, we can see it as a sum of the $k+1$ polynomial matrices $Q_i(y)= P_i [n(y)]^i [d(y)]^{g-i}$, $0\leq i \leq k$, with either $Q_i(y)=P_i=0$ or $\deg Q_i(y) = gD+i(N-D)$. Since the degree of the sum of two polynomials cannot exceed the greatest of the degrees of the considered polynomials, $\deg Q(y)$ cannot be greater than $q$. Notice that if $N=G$ then $gG\geq q=kG+(g-k)D$ and the maximum is realised by $i=k$, while otherwise the maximum is realised by the smallest index $j$ such that $P_j \neq 0$, and $q=(g-j)G+jN$. This means that if $N<G$ and $P_0=0$ then $q<gG$, while $q=gG$ if $N<G$ but $P_0 \neq 0$.
Notice finally that, if $i_1 \neq i_2$, then $Q_{i_1}(y)$ and $Q_{i_2}(y)$ have the same degree if and only if $D=N$. Since
$\deg Q_{i_1}(y) \neq \deg Q_{i_2}(y) \Rightarrow \deg (Q_{i_1}(y)+Q_{i_2}(y))=
\max(\deg Q_{i_1}(y),\deg Q_{i_2}(y))$, $D \neq N$ is a sufficient condition for $\deg Q(y) =q$.
\end{proof}
Lemma \ref{degQ} shows that
$\deg Q(y) \leq q \leq gG$. The next Proposition describes the conditions under which the equality
$\deg Q(y) = gG$ holds.
\begin{proposition}\label{degphipi}
Let $Q(y) =\Phi (P(x))$. It always holds $\deg Q(y) \leq gG$, and $\deg Q(y) < g G$ if and only if one of the following is true:
\begin{enumerate}
\item $N > D$ and $g>k$;
\item $N \leq D$, and there exist a natural number $a \geq 1$ and a polynomial matrix
$\hat{P}(x) \in (\mathbb{F}[x])^{m \times p}$ such that $P(x)=(x-\hat{x})^a \hat{P}(x)$, where
$\hat{x} := n_G d_G^{-1}$ if $N=D=G$ and $\hat{x}:=0_{\mathbb{F}}$ if $N < D =G$.
\end{enumerate}
\end{proposition}
\begin{proof}
Lemma \ref{degQ} guarantees $\deg Q(y) \leq gG$. To complete the proof, there are three possible cases to be analysed.
\begin{itemize}
\item If $G=N>D$, we know from Lemma \ref{degQ} that $\deg Q(y) = q$, and in this case $q=gD+kN-kD$. Therefore,
$\deg Q(y)=gG \Leftrightarrow g=k$.
\item If $N=D=G$ and, we get $q=gG$. Let $Q(y)=\sum_{i=0}^{gG} \Theta_i y^i$: then, $\deg Q(y) < gG
\Leftrightarrow \Theta_{gG} = 0_{(\mathbb{F}[x])^{m \times p}}$. On the other hand $\Theta_{gG}$ is the coefficient of $y^{gG}$ in
$Q(y)=\sum_{i=0}^{g} P_i [n(y)]^i [d(y)]^{g-i}$, so $\Theta_{gG} = d_G^g \sum_{i=0}^g P_i n_G^i d_G^{-i} = d_G^g P(n_G d_G^{-1})$.
Therefore, $\Theta_{gG}$ is zero if and only if every entry of
$P(n_G d_G^{-1})$ is equal to $0_{\mathbb{F}[x]}$, or in other words if and only if
$P(x) = (x-n_G d_G^{-1})^a \hat{P}(x)$ for some $a \geq 1$ and some suitable polynomial matrix $\hat{P}(x)$.
\item If $N<D=G$, recalling the proof of Lemma \ref{degQ} we conclude that $\deg Q(y) < gG$ if and only if $P_0 = 0$, which is
equivalent to $P(x) = x^a \hat{P}(x)$ for a suitable value of $a \geq 1$ and some polynomial matrix $\hat{P}(x)$.
\end{itemize}
\end{proof}
The grade of $Q(y)$ is of course arbitrary, even though it must be greater than or equal to its degree. Since $\deg Q(y) \leq q \leq gG$,
we shall define that the grade of $Q(y)$ is $gG$. This choice has an influence on the infinite elementary divisors of $Q(y)$, as they
are equal to the elementary divisors corresponding to zero of the reversal of $Q(y)$ \emph{taken with respect to its grade}, that is
$\mathrm{Rev}_{(gG)} Q(y)$.
If one is interested in picking a different choice for the grade of $Q(y)$, the following Proposition explains how the infinite elementary divisors change.
\begin{proposition}
Let $P(x) \in (\mathbb{F}[x])^{m \times p}$, with $k=\deg P(x)$. Then the finite elementary divisors and the minimal indices of $P(x)$ do not depend on its grade, while the infinite elementary divisors do. Namely, let $\nu=\mathrm{min}(m,p)$; $x^{g-k} d_1(x),\dots,x^{g-k} d_{\nu}(x)$ are the invariant polynomials of $\mathrm{Rev}_g P(x)$ if and only if $d_1(x),\dots,d_{\nu}(x)$ are the invariant polynomials of $\mathrm{Rev}_k P(x)$, for any choice of $g \geq k$.
\end{proposition}
\begin{proof}
Neither Theorem \ref{smiththeorem} nor the properties of $\ker P(x)$ and $\ker P^T(x)$ depend on the grade, so minimal indices and finite elementary divisors cannot be affected by different choices. Let $S(x)=A(x) \mathrm{Rev}_k P(x) B(x)$ be the Smith form of $\mathrm{Rev}_k P(x)$. We have $\mathrm{Rev}_g P(x)=x^{g-k} \mathrm{Rev}_k P(x)$, which implies that $x^{g-k}S(x)=A(x) \mathrm{Rev}_g P(x) B(x)$. Clearly $d_i(x)|d_j(x) \Leftrightarrow x^{g-k}d_i(x)|x^{g-k}d_j(x)$, and therefore we conclude that $x^{g-k}S(x)$ is the Smith form of $\mathrm{Rev}_g P(x)$.
\end{proof}
Let $\alpha,\beta,\gamma,\delta \in \mathbb{F}$. If $G=1$, $\Phi_{g,\alpha y+\beta,\gamma y+\delta}$ is clearly invertible and its inverse, with a little abuse of notation, is $\Phi_{g,\beta-\delta x,\gamma x-\alpha}(Q(y)) = [\gamma x - \alpha]^g Q(\frac{\beta-\delta x}{\gamma x - \alpha})$
. The most general case is analysed below.
\begin{proposition}\label{nobij}
Let us denote by $\mathbb{F}[x]_g$ the set of the univariate polynomials in $x$ whose degree is less than or equal to $g$. Given $g,n(y),d(y)$, the application $\Phi_{g,n(y),d(y)}:(\mathbb{F}[x]_g)^{m \times p}\rightarrow (\mathbb{F}[y]_{(gG)})^{m \times p}$ is always an injective function, but it is not surjective unless $G = 1$.
\end{proposition}
\begin{proof}
Notice that $\Phi_g$ can be thought as acting componentwise, sending $P(x)_{ij}$ to $Q(y)_{ij}=\Phi_g(P(x)_{ij})$. Thus, it will be sufficient to show that, in the scalar case $\Phi_g : \mathbb{F}[x]_g\rightarrow\mathbb{F}[y]_{(gG)}$, $\Phi_g$ is surjective if and only if $G=1$. This is true because any polynomial that does \emph{not} belong to the set $R_y:=\{a(y)\in \mathbb{F}[y]:a(y)=\sum_{i=0}^{g} a_i [d(y)]^{g-i} [n(y)]^i\}$ cannot belong to the image of $\Phi_g$, and $R_y = \mathbb{F}[y]_{(gG)}$ if and only if $G = 1$.
In fact, if we require that a generic $r \in \mathbb{F}[y]_{(gG)}$ belongs to $R_y$, we find out that the $g+1$ coefficients $a_i$ must satisfy $gG+1$ linear constraints.
To prove injectivity: $\Phi_g(P_1(x))=\Phi_g(P_2(x)) \Rightarrow P_1(x(y))=P_2(x(y)) \Rightarrow P_1(x)=P_2(x)$.
\end{proof}
Proposition \ref{nobij} tells us that, unless $G=1$ (the M\"{o}bius case), not every $Q(y)$ is such that $Q(y)=\Phi(P(x))$ for some $P(x)$.
A couple of additional definitions will turn out to be useful in the following. Let $x_0 \in \mathbb{F}^*$: we define $T_{x_0}$ as the counterimage of $x_0$ under the rational function $x(y)$.
Moreover let $\alpha,\beta \in \mathbb{F}$ be such that $\frac{\alpha}{\beta}=x_0$ and $\alpha$ and $\beta$ are not both zero. For instance, we can pick
$(\alpha,\beta)=(x_0,1_{\mathbb{F}})$ if $x_0 \neq \infty$ and $(\alpha,\beta)=(1_{\mathbb{F}},0_{\mathbb{F}})$ otherwise. Consider the polynomial equation
\begin{equation}\label{fundamental}
\alpha d(y)=\beta n(y).
\end{equation}
Let $S$ be the degree of the polynomial $\alpha d(y)-\beta n(y)$. Equation \eqref{fundamental} cannot have more than $S$ finite roots.
If $S<G$ then we formally say $\infty \in T_{x_0}$.
\begin{remark}\label{discu}
Notice that there are three cases that lead to $S<G$:
\begin{enumerate}
\item $N=D=G$ and $x_0=n_G d_G^{-1}$, so that \eqref{fundamental} becomes $d_G n(y) = n_G d(y)$:
in this case, $S$ is the maximum value of $i$ such that $n_i \neq x_0 d_i$;
\item $N<D=G$ and $x_0 = 0_{\mathbb{F}}$, so that \eqref{fundamental} becomes $n(y) = 0_{\mathbb{F}}$ and $S=N$;
\item $D<N=G$ and $x_0 = \infty$, so that \eqref{fundamental} is $d(y) = 0_{\mathbb{F}}$ and $S=D$.
\end{enumerate}
\end{remark}
We now define the multiplicity $m_0$ of any finite $y_0 \in T_{x_0}$ as the multiplicity of $y_0$ as a solution of the polynomial equation \eqref{fundamental}. If $\infty \in T_{x_0}$, its multiplicity is defined to be equal to $G-S$. Therefore, the sum of the multiplicities of all the (both finite and infinite) elements of $T_{x_0}$ is always equal to $G$, while the sum of the multiplicities of all the finite elements of $T_{x_0}$ is $S$.
The finite elements of $T_{x_0}$ are characterised by the following Proposition.
\begin{proposition}\label{crucial}
Let $y_0 \in \mathbb{F}$ and $x_0 \in \mathbb{F}^*$. Then $y_0 \in T_{x_0}$ if and only if $y_0$ is a solution of \eqref{fundamental} for $\alpha,\beta : x_0=\frac{\alpha}{\beta}$.
Moreover, $\alpha_1 d(y_0)=\beta_1 n(y_0)$ and $\alpha_2 d(y_0)=\beta_2 n(y_0)$ if and only if $\frac{\alpha_1}{\beta_1}=\frac{\alpha_2}{\beta_2}$.
\end{proposition}
\begin{proof}
The definition of $T_{x_0}$ implies the first part of the Proposition. The second part comes from the fact that $x(y)$ is a function.
\end{proof}
Proposition \ref{crucial}, albeit rather obvious, has the following important implication:
\begin{corollary}\label{nocommonroot}
$x_0 \neq x_1 \Leftrightarrow T_{x_0} \cap T_{x_1} = \emptyset$.
Equivalently, $\alpha_1 \beta_2 \neq \alpha_2 \beta_1$ if and only if $[\beta_1 n(y) - \alpha_1 d(y)]$ and $[\beta_2 n(y) - \alpha_2 d(y)]$ $\in \mathbb{F}[y]$ are coprime.
In particular, for any finite $x_0 \in \mathbb{F}$, $\Phi(x-x_0)$ and $d(y)$ are coprime.
\end{corollary}
In order to clarify the latter definitions, let us consider an example.
Let $\mathbb{F}=\mathbb{C}$ and take $n(y)=y^4+y^3-y^2-y+1$, $d(y)=y^4$.
$T_1$ is the set of the solutions of the equation $n(y)=d(y)$, so in this case $T_1=\{-1,1,\infty\}$. Moreover, the multiplicity of $-1$ and $1$ are, respectively, $1$ and $2$;
since $S=3$ and $G=4$, the multiplicity of $\infty$ is by definition $G-S=1$. Within the same example, $T_{\infty}=\{0\}$; $0$ has multiplicity $4$ because it is a root of order $4$ of the equation $d(y)=0$.
\section{Main result}\label{themainresult}
We are now able to state our main Theorem.
\begin{theorem}\label{main}
Given $m,p \in \mathbb{N}_0$ and $n(y),d(y) \in \mathbb{F}[y]$, let $x_0 \in \mathbb{F}^*$ be a characteristic value of $P(x) \in (\mathbb{F}[x])^{m \times p}$, and let $(x-x_0)^{\ell_1}, \dots, (x-x_0)^{\ell_j}$ be the corresponding elementary divisors. Let $g$ be the grade of $P(x)$, define $G=\max(\deg n(y),\deg d(y))$ and let $gG$ be the grade of $Q(y)=\Phi_{g}(P(x)):=[d(y)]^g P(\frac{n(y)}{d(y)}) \in (\mathbb{F}[y])^{m \times p}$. Then for any $y_0 \in T_{x_0}$:
\begin{itemize}
\item $y_0$ is a characteristic value of $Q(y)$;
\item $(y-y_0)^{m_0 \ell_1}, \ \dots, \ (y-y_0)^{m_0 \ell_j}$ are elementary divisors corresponding to $y_0$ for $Q(y)$, where $m_0$ is the multiplicity of $y_0$.
\end{itemize}
Conversely, if $Q(y)=\Phi_g(P(x))$ for some $P(x)$, and if $y_0 \in \mathbb{F}^*$ is a characteristic value of $Q(y)$ with corresponding elementary divisors $(y-y_0)^{\kappa_1}, \ \dots, \ (y-y_0)^{\kappa_j}$:
\begin{itemize}
\item $x_0=\frac{n(y_0)}{d(y_0)}$ is a characteristic value of $P(x)$;
\item $m_0|\kappa_i$ $\forall i \leq j$, where $m_0$ is the multiplicity of $y_0$ as an element of $T_{x_0}$, and $(x-x_0)^{m_0^{-1} \kappa_1}, \ \dots, \ (x-x_0)^{m_0^{-1} \kappa_j}$ are elementary divisors corresponding to $x_0$ for $P(x)$.
\end{itemize}
In addition, the following properties hold:
\begin{itemize}
\item the right minimal indices of $P(x)$ are $\beta_1, \ \dots, \ \beta_s$ if and only if the right minimal indices of $Q(y)$ are $G \beta_1, \ \dots, \ G \beta_s$;
\item the left minimal indices of $P(x)$ are $\gamma_1, \ \dots, \ \gamma_r$ if and only if the left minimal indices of $Q(y)$ are $G \gamma_1, \ \dots, \ G \gamma_r$.
\end{itemize}
\end{theorem}
For any choice of the application $\Phi_g$, Theorem \ref{main} gives a thorough description of the complete eigenstructure of $\Phi_g (P(x))$ with respect to the complete eigenstructure of $P(x)$. Notice that if $x(y)$ is a M\"{o}bius transformation then $m_0 \equiv 1$ and $G=1$, so the complete eigenstructure is unchanged but for the shift from one set of characteristic values to another. This is not the case for more general rational transformations, where other changes do happen.
The structure of the proof of Theorem \ref{main} is the following. First we prove the first part of the Theorem
(the statement on elementary divisors). This is done dividing the statement in three cases:
\begin{enumerate}
\item[1.] $x_0 \in \mathbb{F}$ and $y_0 \in \mathbb{F}$;
\item[2.] $x_0 \in \mathbb{F}$ and $y_0 = \infty$;
\item[3.] $x_0 = \infty$.
\end{enumerate}
We first prove that the statement is true for case 1,
then show that this implies that it is true for case 2. The validity of cases 1 and 2 implies case 3.
Finally, we prove the second part of the Theorem (the statement on minimal indices) with a constructive proof: we
build a minimal basis of $Q(y)$ given a minimal basis of $P(x)$, and vice versa.
\section{Proof of Theorem \ref{main}: elementary divisors}\label{proof1}
The proof relies on the following Lemma, whose statement generalises \cite[Proposition 11.1]{bible}.
The proof of the Lemma and more details are given in Section \ref{app}.
\begin{lemma}\label{lancaster}
Let $P(x) \in (\mathbb{F}[x])^{m \times p}$ and let $Q(x)=A(x)P(x)B(x)$ where $A(x) \in M_m(\mathbb{F}[x])$ and
$B(x) \in M_p(\mathbb{F}[x])$ are both regular, and
suppose that $x_0 \in \mathbb{F}$ is neither a root of $\det A(x) \in \mathbb{F}[x]$ nor a root of
$\det B(x) \in \mathbb{F}[x]$.
Then
$P(x)$ and $Q(x)$ have the same elementary divisors associated with $x_0$.
\end{lemma}
\subsection{Case 1}
Define $\nu:=\min(m,p)$, and
let $P(x)=A(x)T(x)B(x)$ where $A(x)$ and $B(x)$ are unimodular polynomial matrices,
$T(x)=:\mathrm{diag}(\delta_1(x),\dots,\delta_\nu(x))$ is the Smith form of $P(x)$,
and $\delta_i(x)$ are its invariant polynomials.
Let now $\hat{Q}(y):=\Phi(A(x))\Phi(T(x))\Phi(B(x))$. Clearly, $\hat{Q}(y)$ and $Q(y)$ differ only for a
multiplicative factor $[d(y)]^{\lambda}$, $\lambda \in \mathbb{N}$; moreover,
both $\det \Phi(A(x))$ and
$\det \Phi(B(x))$ are nonzero whenever $d(y)\neq 0_{\mathbb{F}}$.
Notice that, if $x_0$ is finite, then for any $y_0 \in T_{x_0}$ there must hold $d(y_0)\neq 0_{\mathbb{F}}$
(Corollary \ref{nocommonroot}). Therefore, Lemma \ref{lancaster} implies that $Q(y)$, $\hat{Q}(y)$ and $S(y)$ have
the same elementary divisors corresponding to $y_0$.
Unfortunately, $\Phi(T(x))$ may not be the Smith form of $\hat{Q}(y)$, because neither $\Phi(A(x))$ nor $\Phi(B(x))$
are necessarily unimodular and
also because $\Phi(\delta_i(x))$ may not be monic. Nevertheless, it has the form
$\mathrm{diag}([d(y)]^{k_1}\hat{\delta}_1(y), \dots, [d(y)]^{k_{\nu}}\hat{\delta}_{\nu}(y))$, where
$k_1 \geq k_2 \geq \dots \geq k_{\nu}$ and
$\hat{\delta_i}(y):=\Phi(\delta_i(x))$. From Corollary \ref{nocommonroot}, $\hat{\delta}_i(y)$ and $d(y)$ cannot share common roots.
To reduce $S(y)$ into a Smith form, we proceed by steps working on $2 \times 2$ principal submatrices.
In each step, we consider the submatrix $\left[\begin{smallmatrix} [d(y)]^{\gamma}
\hat{\delta}_i(y) & 0\\0 &
[d(y)]^{\phi}\hat{\delta}_{j}(y)\end{smallmatrix}\right]$, where $\gamma:=k_i$ and $\phi:=k_j$, with $i<j$. If
$\gamma=\phi$, then do nothing;
if $\gamma > \phi$, premultiply the submatrix by $\left[\begin{smallmatrix} 1_{\mathbb{F}} &
1_{\mathbb{F}}\\-b(y)q(y) & 1_{\mathbb{F}}-b(y)q(y)\end{smallmatrix}\right]$
and postmultiply it by $\left[\begin{smallmatrix} a(y) &
-q(y)\\b(y) & [d(y)]^{\gamma-\phi}\end{smallmatrix}\right]$,
where $q(y)=\hat{\delta}_{j}(y)/\hat{\delta}_i(y)$ while $a(y)$
and $b(y)$ are such that $a(y) [d(y)]^{\gamma}
\hat{\delta}_i(y) + b(y) [d(y)]^{\phi} \hat{\delta}_{j}(y) =
[d(y)]^{\phi} \hat{\delta}_i(y)$; the existence of two such
polynomials is guaranteed by Bezout's lemma, since $[d(y)]^{\phi} \hat{\delta}_i(y)$ is
the greatest common divisor of $[d(y)]^{\gamma} \hat{\delta}_i(y)$
and $[d(y)]^{\phi}\hat{\delta}_{j}(y)$. It is easy to check that both matrices are
unimodular, and that the result of the matrix multiplications
is $\left[\begin{smallmatrix} [d(y)]^{\phi}\hat{\delta}_i(y) & 0\\0 &
[d(y)]^{\gamma} \hat{\delta}_{j}(y)\end{smallmatrix}\right]$.
Hence, by subsequent applications of this algorithm and after having defined a unimodular diagonal matrix
$\Delta \in \mathbb{F}^{\nu \times \nu}$ chosen in such a way that the invariant polynomials of $S(y)$ are monic, it is possible
to conclude that the Smith form of
$\Phi(T(x))$ is either $S(y):=\Delta \cdot \hat{S}(y)$ or $S(y):= \hat{S}(y)\cdot \Delta$ (whichever of the two products makes sense,
depending on whether $m \leq p$ or not), where
$$\hat{S}(y)=\textrm{diag}(
[d(y)]^{k_m}\hat{\delta}_1(y),\dots,[d(y)]^{k_1}\hat{\delta}_m(y)).$$
Thus, the $i$th invariant polynomial of $P(x)$ has a root of multiplicity $\ell_i$ at $x_0$ if and only if the $i$th invariant
polynomial of $\hat{Q}(y)$ has a root of multiplicity $m_0 \ell_i$ at $y_0 \in T_{x_0}$.
\subsection{Case 2}
By definition, the infinite elementary divisors for a given polynomial matrix are the elementary divisors corresponding
to zero of the reversal of such polynomial matrix. Therefore, in order to prove Theorem \ref{main} for the case of $y_0=\infty$, we
have to analyse the polynomial matrix $Z(y):=\mathrm{Rev}_{(gG)} Q(y) = y^{gG} [d(y^{-1})]^g P(x(y^{-1}))$, and find out what its
relation to $P(x)$ is, with particular emphasis to its elementary divisors corresponding to $y_0=0_{\mathbb{F}}$.
Recalling Remark \ref{discu},
notice that there are two distinct subcases for which $\infty \in T_{x_0}$ for a finite
$x_0 \in \mathbb{F}$. We will consider them separately.
\subsubsection{Subcase 2.1: $N=D=G$, $x_0=n_G d_G^{-1}$}
We get $x(y^{-1})=\frac{\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)}$ and $y^G d(y^{-1})=\mathrm{Rev} d(y)$; therefore
$Z(y)=[\mathrm{Rev} d(y)]^g P(\frac{\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)})$. This means that we can prove analogous results for
$Z(y)$ just by considering this time the new rational transformation $y \rightarrow x=\frac{\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)}$.
From Remark \ref{discu}, $0_{\mathbb{F}}$ is a root of multiplicity $G-S$ for the equation
$\mathrm{Rev} n(y)=x_0 \mathrm{Rev} d(y)$; moreover, since we took the reversal with respect to the degree
(or also because of Corollary \ref{nocommonroot}), $0_{\mathbb{F}}$ cannot be a root of $\mathrm{Rev}d(y)$.
Therefore, following the proof given above, one can state that $P(x)$ has $(x-x_0)^{\ell_1}, \dots, (x-x_0)^{\ell_j}$ as
elementary divisors corresponding to $x_0$ if and only if $Z(y)$ has the $j$ elementary divisors
$y^{(G-S) \ell_1}, \dots, y^{(G-S)\ell_j}$ corresponding to $0_{\mathbb{F}}$. The thesis follows immediately.
\subsubsection{Subcase 2.2: $N<D=G$, $x_0 = 0_{\mathbb{F}}$}
This time, we can write $x(y^{-1})=\frac{y^{G-N}\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)}$ and $Z(y)=[\mathrm{Rev} d(y)]^g$ $P(\frac{y^{G-N}\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)})$. It is therefore sufficient to consider the transformation $y \rightarrow x=y^{G-N}\frac{\mathrm{Rev} n(y)}{\mathrm{Rev} d(y)}$.
In fact, notice that $0_{\mathbb{F}}$ is a solution of multiplicity $G-N$ for the equation $y^{G-N}\mathrm{Rev} n(y)=0$ ($0_{\mathbb{F}}$ is neither a root of $\mathrm{Rev} n(y)$ nor a root of $\mathrm{Rev} d(y)$, because
$\mathrm{Rev} n(0_{\mathbb{F}})=n_N \neq 0_{\mathbb{F}}$ and $\mathrm{Rev} d(0_{\mathbb{F}})=d_D \neq 0_{\mathbb{F}}$). Thus, $P(x)$ has the $j$ elementary divisors $x^\ell_1, \dots, x^\ell_j$ corresponding to $0_{\mathbb{F}}$ if and only if $Z(y)$ has the $j$ elementary divisors $y^{(G-N) \ell_1}, \dots, y^{(G-N)\ell_j}$ corresponding to $0_{\mathbb{F}}$, and the thesis follows.
\subsection{Case 3}
By definition, the infinite elementary divisors of $P(x)$ are the elementary divisors corresponding to the characteristic value $0_{\mathbb{F}}$ for $R(x):=\mathrm{Rev}_g P(x)=x^g P(x^{-1})$. But let $\Psi_{g,n(y),d(y)}=\Phi_{g,d(y),n(y)}$ and $U(y)=\Psi_g(R(x))$, that is to say $U(y) = [n(y)]^g R(\frac{d(y)}{n(y)})$. A simple calculation gives
\[U(y)=[n(y)]^g[\frac{d(y)}{n(y)}]^g P([\frac{d(y)}{n(y)}]^{-1})=[d(y)]^g P(\frac{n(y)}{d(y)})=\Phi_g(P(y))=Q(y).\]
One can therefore follow the proof as in the previous Subsections, but starting from $R(x)$ and using a different transformation (notice
that the equation $d(y)=0_{\mathbb{F}}$ defines both
$T_{\infty}$ for the old transformation and $T_{0_{\mathbb{F}}}$ for the new transformation).
\section{Proof of Theorem \ref{main}: minimal indices}\label{proof2}
We shall only prove the theorem for right minimal indices. The proof for left minimal indices follows from the proof for right minimal indices and from the fact that $\Phi$ and the operation of transposition commute, that is $\Phi_g(P^T(x))=(\Phi_g(P(x)))^T$ $\forall$ $P(x) \in (\mathbb{F}[x])^{m \times p}$.
\subsection{$\Rightarrow$}
Let $\dim \ker P(x)=s$, and $V(x)=[v_1(x),\dots,v_s(x)]$ be a minimal basis for $\ker P(x)$, with minimal indices
$\beta_i:=\deg v_i$ $\forall i=1,\dots,s$ and order $B:=\sum_{i=1}^s \beta_i$. For each value of $i$ let us define
$w_i(y): = \Phi_{\beta_i}(v_i(x))$; we also define $W(y):=[w_1(y),\dots,w_s(y)]$. Clearly $\deg w_i(y) = G \beta_i$.
Suppose in fact $\deg w_i(y) \neq G \beta_i$; applying Proposition \ref{degphipi} (in the case $g=k=\beta_i$), this would imply that there exists some $x_0 \in \mathbb{F}$ and some polynomial vector $u(x) \in (\mathbb{F}[x])^p$ such that $v_i(x)=(x-x_0) u(x)$. Hence, $[v_1(x),\dots,(x-x_0)^{-1}v_i(x),\dots,v_s(x)]$ would be a polynomial basis of order $B-1$ for $\ker P(x)$, leading to a contradiction. In order to prove that $W(y)$ is a minimal basis for $\ker Q(x)$ we must show that it is a basis and that it is minimal.
Clearly $w_i(y)$ lies in $\ker Q(y)$ for all $i$. In fact, $P(x) v_i (x) = 0$ implies that $Q(y) w_i (y) = 0$.
So it is sufficient to show that $W(y)$, considered as an element of $(\mathbb{F}(x))^{p \times s}$, has rank $s$. Notice that
$W(y)=V(x(y))\cdot\mathrm{diag}([d(y)]^{\beta_1},\dots,[d(y)]^{\beta_s})$. A well-known property of the rank is that, if
$A_1=A_2 A_3$ and $A_3$ is square and regular, then $\mathrm{rk}(A_1)=\mathrm{rk}(A_2)$. Therefore
$\mathrm{rk}(W(y))=\mathrm{rk}(V(x(y)$), because the diagonal matrix above is regular. Let $\hat{V}(x)$ be some regular
$s \times s$ submatrix of $V(x)$, which exists because $\mathrm{rk}(V(x))=s$.
By hypothesis, $\det(\hat{V}(x)) \neq 0_{\mathbb{F}[x]}$, which implies $\det(\hat{V}(x(y)))\neq 0_{\mathbb{F}(y)}$. Hence $s=\mathrm{rk}(V(x(y)))=\mathrm{rk}(W(y))$. Then $W(y)$ is a basis.
In order to prove that it is minimal, let us introduce the following lemma whose proof can be found in \cite{forneyjr}.
\begin{lemma}\label{forneyjr}
Let $\mathcal{V}$ be a vector subspace of $\mathbb{F}(x)^p$, with $\dim \mathcal{V} = s$. Let $H=[h_1,\dots,h_s]$ be a polynomial
basis of order $A$ for $\mathcal{V}$ and define $\xi_i$, $i=1,\dots,\left(\begin{smallmatrix}
p\\
s\\
\end{smallmatrix}\right)$ to be the $s \times s$ minors (i.e. determinants of $s \times s$ submatrices) of $H$. Then the following statements are equivalent:
\begin{itemize}
\item $H$ is a minimal basis for $\mathcal{V}$
\item The following conditions are both true: (a) $\mathrm{GCD}(\xi_1, \dots, \xi_r)=1_{\mathbb{F}[x]}$ and (b) $\max_i$ $\deg \xi_i = A$.
\end{itemize}
\end{lemma}
So let $\xi_i(y)$ be the $s \times s$ minors of $W(y)$. We shall prove that (a) their GCD is $1_{\mathbb{F}[y]}$ and (b) their maximal degree is $G B = G \sum_{i=1}^s \beta_i$. By Lemma \ref{forneyjr}, these two conditions imply that $W(y)$ is minimal. Recall that $w_i(y) = \Phi_{\beta_i} (v_i(x))$, that is to say $w_i (y)=[d(y)]^{\beta_i} v_i(x(y))$. Any $s \times s$ submatrix of $W(y)$ is therefore obtained from the corresponding $s \times s$ submatrix of $V(x)$ by applying the substitution $x=x(y)$ and then multiplying the $i$th column by $[d(y)]^{\beta_i}$ for $i=1,\dots,s$. Let us call $\zeta_i(x)$ the $s \times s$ minors of $V(x)$. From the properties of determinants we obtain the relation $\xi_i (y)= \left(\prod_{i=1}^s [d(y)]^{\beta_i} \right) \zeta_i (x(y)) = [d(y)^B] \zeta_i (x(y)) = \Phi_B (\zeta_i (x))$.
Now for each $i$ let $\gamma_i := \deg \zeta_i (x)$ and $\delta_i :=\deg \xi_i (y) \leq \max_{j \leq \gamma_i}(N j-D j) + DB$ where the maximum is taken over those values of $j$ such that the $j$th coefficient of $\xi_i(y)$ is nonzero (Lemma \ref{degQ}). There are two cases.
If $N \leq D=G$, $\delta_i \leq GB$, and applying Proposition \ref{degphipi} (with $g=B$), the inequality holds if and only if
if $(x-\hat{x})|\zeta_i(x)$, where $\hat{x}=0_{\mathbb{F}}$ if $N<D$ and $\hat{x}=n_G d_G^{-1}$ if $N=D$;
notice that there must be at least one value of $i$ for which $\delta_i = GB$, otherwise $(x-\hat{x})$ would be a common factor of all
the $\zeta_i(x)$, which is not possible because of Lemma \ref{forneyjr}.
Finally, if $D<N=G$, $\delta_i = \gamma_i G + (B - \gamma_i) D$. Since $V(x)$ is minimal we have $\max_i(\gamma_i)=B$,
which implies that also in this case $\max_i(\delta_i)=GB$. This proves condition (b).
Notice moreover that $\xi_i (y)=\Phi_B(\zeta_i(x))=[d(y)]^{B-\gamma_i}\Phi_{\gamma_i}(\zeta_i(x))$, where the first and the second factor are coprime (because of Corollary \ref{nocommonroot}). Let us prove the following Lemma.
\begin{lemma}\label{phiofgcd}
Let $p, q, r \in \mathbb{F}[x]$ with $r$ monic. Then, $\mathrm{GCD}_{\mathbb{F}[x]}(p,q)=r$ if and only if $\mathrm{GCD}_{\mathbb{F}[y]}(\Phi_{\deg p}(p),\Phi_{\deg q}(q))=\kappa \cdot \Phi_{\deg r}(r)$, where $\kappa \in \mathbb{F}$ is such that $\kappa \cdot \Phi_{\deg r}(r)$ is monic.
\end{lemma}
\begin{proof}
Let $\alpha, \beta$ be two suitable elements of $\mathbb{F}$ and let us write the prime factor decompositions
$p=\alpha \cdot \prod (x - p_i)^{\pi_i}$, $q=\beta \cdot \prod (x - q_i)^{\theta_i}$, $r=\prod (x - r_i)^{\rho_i}$. Of course we have
that $(x-r_i)^{\rho_i}|r$ if and only if $(x-p_i)^{\pi_i}|p$, $(x-q_i)^{\theta_i}|q$ and $\rho_i = \min (\pi_i,\theta_i)$.
We get $\Phi_{\deg p}(p) = \alpha \cdot \prod (n(y) - p_i d(y))^{\pi_i}$,
$\Phi_{\deg q}(q) = \beta \cdot \prod (n(y) - q_i d(y))^{\theta_i}$ and
$\Phi_{\deg r}(r) = \prod (n(y) - r_i d(y))^{\rho_i}$. The thesis follows by invoking Corollary \ref{nocommonroot}.
\end{proof}
Lemma \ref{phiofgcd} implies condition (a). This follows from the equation $\mathrm{GCD}_i(\xi_i(y))=\mathrm{GCD}_i([d(y)]^{B-\gamma_i}) \cdot \mathrm{GCD}_i (\Phi_{\gamma_i}(\zeta_i(x)))=1_{\mathbb{F}[y]} \cdot 1_{\mathbb{F}[y]}$, where the first $1_{\mathbb{F}[y]}$ comes from the fact that $\max_i(\gamma_i)=B$, while the second $1_{\mathbb{F}[y]}$ comes by applying the previous Lemma to $\mathrm{GCD}(\xi_1(y),\dots,\xi_s(y))=\mathrm{GCD}(\mathrm{GCD}(\dots\mathrm{GCD}(\xi_{2}(y),\xi_{1}(y))\dots))$ and from the identity $\Phi_0(1_{\mathbb{F}[x]})=1_{\mathbb{F}[y]}$.
\subsection{$\Leftarrow$}
To complete the proof, suppose now that $Q(y)=\Phi_g(P(x))$ for some $P(x) \in \mathbb{F}[x]$ and that $\hat{W}(y)$ is a minimal basis for $\ker Q(y)$, with minimal indices $\epsilon_1 \leq \dots \leq \epsilon_s$. The other implication that we proved in the previous subsection implies that $G|\epsilon_i$ $\forall$ $i$, so define $\beta_i=\frac{\epsilon_i}{G}$. Suppose that there exists a minimal basis $\hat{V}(x)=(\hat{v}_1(x),\dots,\hat{v}_s(x))$ for $\ker P(x)$; suppose moreover that an index $i_0 \in \{1,\dots,s\}$ exists such that $\deg \hat{v}_{i_0} \neq \beta_{i_0}$. Applying the reverse implication, this would imply that there is a minimal basis $\tilde{W}(y)=(\tilde{w}_1(y),\dots,\tilde{w}_s(y))$ for $\ker Q(y)$ whose $i_0$th right minimal index is not equal to $\epsilon_{i_0}$. This is absurd because every minimal basis has the same minimal indices.
\section{Extension to more relaxed hypotheses}\label{comments}
For the sake of convenience, we have so far assumed that the field $\mathbb{F}$ is algebraically closed.
This unnecessary hypothesis can be dropped. To see it, assume that $\mathbb{F}$ is not algebraically closed and
let $\mathbb{K}$ be the algebraic closure of $\mathbb{F}$.
Then $(\mathbb{F}[x])^{m \times p} \subseteq (\mathbb{K}[x])^{m \times p}$, so we can use Theorem \ref{main} to identify the Smith forms
of $P(x)$ and $Q(y)=\Phi_g(P(x))$ over the polynomial rings $\mathbb{K}[x]$ and $\mathbb{K}[y]$. We can then join back elementary
divisors in $\mathbb{K}[x]$ and $\mathbb{K}[y]$ to form elementary divisors in $\mathbb{F}[x]$ and $\mathbb{F}[y]$. Of course, in this
case an elementary divisor is no more necessarily associated with a characteristic value in $\mathbb{F}$. For instance, if
$\mathbb{F}=\mathbb{Q}$, then the elementary divisor $x^2+2$ is not associated with any rational characteristic value, but if we consider
the field of complex algebraic numbers $\mathbb{K}=\overline{\mathbb{Q}}$ then we can split it as $(x-\sqrt{2}i)(x+\sqrt{2}i)$ and
associate it to the characteristic values $\pm \sqrt{2}i$.
Similarly, the other results (e.g., Lemma \ref{phiofgcd}) that use the algebraic closure of $\mathbb{F}$ can be straightforwardly
extended to a generic field
$\mathbb{F}$ via an immersion into its algebraic closure $\mathbb{K}$.
\section{Proof of Lemma \ref{lancaster}}\label{app}
Let $P(x) \in \mathbb{F}^{m \times p}[x]$. If $U(x):=[u_1(x), \ \dots, \ u_s(x)]$ is a minimal basis for $\ker P(x)$,
we define $\ker_{x_0} P(x):=\mathrm{span} \left(\{u_1(x_0), \ \dots, \ u_s(x_0)\}\right) \subseteq \mathbb{F}^p$.
In general $\ker_{x_0} P(x)$ is a subset of $\ker P(x_0)$. It is
a proper subset when $x_0$ is a characteristic value of $P(x)$, as is illustrated by the following example: let
$\mathbb{F}=\mathbb{C}$ and
\[
P(x)=\left[\begin{array}{cccc}
x & 1 & 0 & 0\\
0 & x & 1 & 0\\
0 & 0 & 0 & 0\\
0 & 0 & 0 & x\\
\end{array}\right].
\]
Evaluating the
polynomial at $0$, we get
$\ker P(0)=\mathrm{span}(\{[1,0,0,0]^T,[0,0,0,1]^T\})$.
On the other hand, a minimal basis for $\ker P(x)$ is $[1,-x,x^2,0]^T$, so $\ker_{0} P(x)=\mathrm{span}(\{[1,0,0,0]^T\})$.
We need now to slightly modify a definition given in \cite{bible} in order to extend it to
the case of singular and/or rectangular polynomial
matrices.
A polynomial vector $v(x) \in (\mathbb{F}[x])^{p}$ is called a \emph{root polynomial} of order $\ell$ corresponding
to $x_0$ for $P(x)$ if the following conditions are met:
\begin{enumerate}
\item $x_0$ is a zero of order $\ell$ for $P(x)v(x)$;
\item $v(x_0) \not\in \ker_{x_0} P(x)$.
\end{enumerate}
Observe that $v(x_0) \in \ker_{x_0} P(x) \Leftrightarrow
\exists \ w(x) \in \ker P(x) \subseteq (\mathbb{F}(x))^p : w(x_0)=v(x_0)$.
In fact, let $w(x)=U(x)c(x)$ for some $c(x) \in (\mathbb{F}(x))^s$ and $w(x_0)=v(x_0)$:
then $v(x_0)=U(x_0)c(x_0) \in \ker_{x_0} P(x)$. Conversely, write $v(x_0)=U(x_0)c$ for some $c \in \mathbb{F}^s$ and notice that
$U(x)c \in \ker P(x)$. Hence, condition 2. implies $v(x) \not\in \ker P(x)$.
In \cite[Proposition 1.11]{bible} it is shown that given three \emph{regular}
polynomial matrices $P(x), A(x), B(x) \in M_n(x)$, and if $x_0$ is neither a root of $\det A(x)$ nor a root of $\det B(x)$, then $v(x)$
is a root polynomial of order $\ell$ for $A(x)P(x)B(x)$ corresponding to $x_0$ if and only if $B(x)v(x)$ is a root polynomial of order
$\ell$ corresponding to $x_0$ for $P(x)$. The next Proposition generalises this result.
\begin{proposition}\label{PandAPB}
Let $P(x) \in \mathbb{F}^{m \times p}[x]$, $A(x) \in M_m(\mathbb{F}[x])$ and $B(x) \in M_p(\mathbb{F}[x])$. Suppose that both $A(x_0)$
and $B(x_0)$, with $x_0 \in \mathbb{F}$, are full rank matrices. Then $v(x)$
is a root polynomial of order $\ell$ corresponding to $x_0$ for $A(x)P(x)B(x)$ if and only if $B(x)v(x)$ is a root polynomial of order
$\ell$ corresponding to $x_0$ for $P(x)$.
\end{proposition}
\begin{proof}
Notice that if $A(x_0)$ and $B(x_0)$ are full rank then $A(x)$ and $B(x)$ are regular.
In \cite{bible}, a root polynomial
is defined for regular square polynomial matrices, so that condition 2. reduces to $v(x_0) \neq 0$. Nevertheless, the proof given
in \cite[Proposition 1.11]{bible} for condition 1. does not actually use the regularity of $P(x)$, and it is therefore
still valid when $P(x)$ is not a regular square polynomial matrix. To complete the proof:
$v(x_0) \in \ker_{x_0} A(x)P(x)B(x) \Leftrightarrow \exists w_1(x) \in \ker A(x)P(x)B(x) : w_1(x_0)=v(x_0) \Leftrightarrow \exists w_2(x)
\in \ker P(x) : w_2(x_0)=B(x_0)v(x_0) \Leftrightarrow B(x_0)v(x_0) \in \ker_{x_0} P(x)$. To build $w_2(x)$ from $w_1(x)$, simply put
$w_2(x)=B(x)w_1(x)$ and use the fact that $A(x)$ is regular. To build $w_1(x)$ from $w_2(x)$, let $(B(x))^{-1}$ be the inverse matrix
(over $\mathbb{F}(x)$) of $B(x)$, which exists because $B(x)$ is regular; then, put $w_1(x)= (B(x))^{-1} w_2(x)$.
\end{proof}
Let $v_1(x),\dots,v_s(x)$ be root polynomials corresponding to $x_0$ of orders $\ell_1 \leq \dots \leq \ell_s$.
We call them a \emph{maximal set of $x_0$-independent root polynomials} if:
\begin{enumerate}
\item they are $x_0$-independent, i.e. $v_1(x_0),\dots,v_s(x_0)$ are
linearly independent;
\item no $(s+1)$-uple of
$x_0$-independent root polynomials corresponding to $x_0$ exists;
\item there are no root polynomials corresponding to $x_0$ of order $\ell>\ell_s$;
\item for all $j=1,\dots,s-1$, there does not exist a root polynomial $\hat{v}_{j}(x)$ of order $\hat{\ell}_{j} > \ell_j$ such that
$\hat{v}_j(x),v_{j+1}(x),\dots,v_s(x)$ are $x_0$-independent.
\end{enumerate}
As long as $\det B(x_0)$ and $\det A(x_0)$ are nonzero, it is easy to check that
$v_1(x),\dots,v_s(x)$ are a maximal set of $x_0$-independent
root polynomials for
$A(x)P(x)B(x)$ if and only if $B(x)v_1(x),\dots,B(x)v_s(x)$ are
a maximal set of $x_0$-independent
root polynomials for $P(x)$.
The next Proposition completes the proof of Lemma \ref{lancaster}.
\begin{proposition}\label{rootandsmith}
$P(x) \in (\mathbb{F}[x])^{m \times p}$ has a maximal set of $x_0$-independent
root polynomials, of order $\ell_1, \dots, \ell_s$, if
and only
if $(x-x_0)^{\ell_1},\dots,(x-x_0)^{\ell_s}$ are the elementary divisors of $P(x)$ associated with $x_0$.
\end{proposition}
\begin{proof}
Let $S(x)$ be the Smith form of $P(x)$, and recall that the inverse of a unimodular polynomial matrix
is still a unimodular
polynomial matrix \cite{bible}. Thus, in view of Proposition \ref{PandAPB} and Theorem \ref{smiththeorem},
it suffices to prove the thesis for $S(x)$.
If $S(x)$ is the zero matrix,
it has neither a root polynomial nor an elementary divisor, so there is nothing to prove. Otherwise, let $\nu$ be the maximal value of
$i$ such that $(S(x))_{ii}\neq 0_{\mathbb{F}[x]}$ and for $j=1,\dots,p$ let $e_j \in (\mathbb{F}[x])^p$ be the
polynomial vector such
that $(e_j)_i=\delta_{ij}$. If $\nu < p$, $[e_{\nu+1},\dots,e_p]$ is a minimal basis for $\ker S(x)$ and,
being of order $0$, also
for $\ker_{x_0} S(x)$. Suppose that $v_1(x),\dots,v_s(x)$ is a maximal set of $x_0$-independent root polynomials for $S(x)$.
Let $k \leq
\nu$ be the smallest index such that $(v_s(x_0))_k \neq 0_{\mathbb{F}}$: there must exist such an index because
$v_s(x_0) \not\in \ker_{x_0} S(x)$.
Let $(S(x))_{kk}=(x-x_0)^{\mu} \theta(x)$, with $\theta(x_0)\neq 0_{\mathbb{F}}$. We get $(S(x)v_s(x))_k =
(x-x_0)^{\mu} \theta(x) (v_s (x))_k$, so $\mu \geq \ell_s$. Actually, $\mu=\ell_s$, or $e_k$ would be a root
polynomial of order greater than $\ell_s$, which is absurd.
Then let $k'$ be the largest index not equal to $k$ and
such that $(v_{s-1}(x_0))_{k'} \neq 0_{\mathbb{F}}$ (if such and index does not exist, then $v_{s-1}(x_0)$
is, up to a vector lying in $\ker_{x_0} S(x)$, a multiple of $e_k$ and thus
$\ell_{s-1}=\ell_s$: in this case, replace without any loss of generality $v_{s-1}(x)$ by
a suitable linear combination of $v_{s-1}(x)$ and $v_s(x)$). Following an argument similar as above, we can show that
$(S(x))_{k'k'}=(x-x_0)^{\ell_{s-1}}\hat{\theta}(x)$, $\hat{\theta}(x_0) \neq 0_{\mathbb{F}}$.
We repeat the process until we find all the $s$ sought elementary divisors. There cannot be more,
otherwise $\dim \ker S(x_0) - \dim \ker_{x_0} S(x)>s$ and it would be possible
to find an $(s+1)$-uple of $x_0$-independent root polynomials.
Conversely, it is easy to check that $e_{\nu-s+1},\dots,e_{\nu}$ are a maximal set of $x_0$-independent root polynomials.
\end{proof}
\begin{remark}
Root polynomials carry all the information on Jordan chains \cite{bible}. Let
$v(x)=\sum_{i=0}^{\ell-1}(x-x_0)^i v_i$, $v_i \in \mathbb{F}^m$, be a root polynomial of order $\ell$ corresponding to $x_0$ for $P(x)$.
It is possible to prove that then $w(y)=\sum_{i=0}^{\ell-1} [d(y)]^{\ell-1-i}[n(y)-x_0 d(y)]^i v_i$ is a root polynomial of
order $m_0 \ell$ corresponding to $y_0$ for $Q(y)$. The latter formula
relates the Jordan chains of $Q(y)$ at $y_0$ to the Jordan chains of $P(x)$ at $x_0$.
\end{remark}
\section{Conclusions}
We have shown that if $P(x)$ and $Q(y)$ are polynomial matrices whose entries belong to the ring of univariate
polynomials in $x$ (resp. $y$) with coefficients in any field, and if $P(x)$ and $Q(y)$ are related by a rational transformation $x(y)$,
then the complete eigenstructures of $Q(y)$ and $P(x)$ are simply related.
\section{Acknowledgements}
I would like to thank N. Mackey for her talk \cite{iciam} on M\"{o}bius transformations that was one of the sources of inspiration for this generalisation on generic rational functions (the other one is \cite{dicksonpal}).
I am also grateful to N. Mackey and D. S. Mackey for the subsequent discussion, and to D. A. Bini and L. Gemignani for useful comments.
Finally, I am indebted with an anonymous referee for valuable suggestions and remarks that helped improving the paper.
|
\section{Introduction}
Congestion games, introduced by Rosenthal \cite{Rosenthal}, form a very natural model for studying many real-life strategic settings: Traffic problems, load balancing, routing, network planning, facility locations and more. In a congestion game players must choose some subset of resources from a given set of resources (e.g., a subset of edges leading from the Source to the Target on a graph). The congestion of a resource is a function of the number of players choosing it and each player seeks to minimize his total congestion accross all chosen resources.
In many modeling instances players and the decision making entity have been thought of as one and the same. However, in a variety of settings this may not be the case.
Consider, for example, a traffic routing game where each driver chooses his route in order
to minimize his travel time, while accounting for congestion along the route caused by other drivers.
Now, in many cases drivers are actually employees
in shipping firms, and in fact it is in the interest of the shipping firm to minimize the total travel time of its fleet.
Similarly, routers in a communication network participate in a congestion game. However, as various routers may belong to the same ISP we are again in a setting where coalitions naturally form. This motivated Fotakis et al. \cite{Fotakis} and Hayrapetyan et al. \cite{Hayra} to introduce the notion of Coalitional Congestion Games (CCG). In a CCG we think about the coalitions as players and each coalition maximizes its total utility.
The coalitional congestion game inherits its structure from the original game,
once the coalitions of players from the original congestion games
(now, becoming the players of the coalitional congestion game) have been identified.
The most notable property of congestion games is that they have posses pure Nash equilibrium. This has been shown by Rosenthal in \cite{Rosenthal}. Later, Monderer and Shapley \cite{Monderer Shapley 96}
formally introduce potential games and show the equivalence of these two classes. The fact that potential games posses a pure Nash equilibrium is straightforward. Unfortunately, the statement that a CCG is a potential game or that it possesses a pure Nash equilibrium is generally false.
In this paper we investigate conditions under which this statement is true. We focus on a subset of congestion games called simple congestion games, where each player is restricted to choose a single resource.
Our main contributions are:
\begin{enumerate}
\item
Whenever each coalition contains at most two players the CCG induced from a simple congestion game possesses a pure-strategy Nash equilibrium (Theorem \ref{Pairs Pure NE}).
\item
If some coalition contains three players, then there may not exist a pure-strategy
Nash equilibrium (Example \ref{No NE}).
\item
If a the congestion game is not simple then there may not exist
a pure-strategy Nash equilibrium (Example \ref{not SCG underlying});
and
\item
Suppose there exists at least one singleton coalition and at least one coalition composed of two players, then a coalitional congestion game induced from
a simple congestion game is a potential game if and only if cost functions are linear (Theorem \ref{linear costs}).
\end{enumerate}
Our results extend and complement the results in Fotakis et al. \cite{Fotakis} and Hayrapetyan et al. \cite{Hayra}.
For example, Fotakis et al. \cite{Fotakis} show that if the resource cost
functions are linear then the coalitional congestion game is a potential
game. We show that, with some additional mild conditions on the
partition structure, this is also a necessary condition.
Hayrapetyan \cite{Hayra} shows that if the underlying congestion
game is simple and costs are weakly convex then the game possesses a pure
Nash equilibrium. We demonstrate additional settings where this holds.
Section 2 provides a model of a coalitional congestions games, section 3 discusses the conditions for the existence of a pure Nash equilibrium in such games and section 4 discusses the conditions for the existence of a potential function.
All proofs are relegated to an appendix.
\section{Model}
Let $G=\{N,S,U\}$ be a non-cooperative game in strategic form. Let
$C=\{C_1, \ldots, C_{n^c}\}$ be a Partition of $N$ into $n^c$
nonempty sets. Hence: $\cup_{k=1}^{n^c} C_k = N$ and $C_k \cap C_l
= \emptyset \;\; \forall k \neq l \in [1, \ldots n^c]$.
The game $G$ and the partition $C$ form a Coalitional
Non-Cooperative (CNC) Game $G^C = \{N^C,S^c,U^c\}$ defined as
follows:\begin{itemize} \item $N^C$ is the set of agents which are
the elements of $C$. \item The strategy space is $S^c =
\{S^c_k\}_{k \in C}$ where $S^c_k = \times_{i \in C_k}S_i$.
\end{itemize}
Note that $\times S_k = \times_{k \in C}\times_{i \in C_k}S_{k,i}$
is isomorphic to $S=\times_{i=1}^n S_i$, since we only changed the
order of the coordinates. Thus, we can look on $s^c$ as a vector
in $S$.
\begin{itemize}
\item The utility function is defined as follows: $\forall s^c \in
S^c \;\; U^c_k(s^c) = \sum_{ i \in C_k}U_i(s^c)$ and $U^c
=\{U^c_k\}_{k \in C}$.
\end{itemize}
In the context of $G^C$, $G$ is the \emph{Underlying Game} and a
player in $G$ is referred to as a \emph{sub agent}. As always, a
Pure Nash Equilibrium of the game $G^C$ is a strategy profile
$s \in S^C$ such that $\forall k \in N^C$, $U^c_k(s) \geq
U^c_k(s_{-k},t_k) \;\;\forall t_k \in S^c_k$. $NE(G^C)$ the
(possibly empty) set of Pure Nash equilibrium strategy profiles in $G^C$.
A congestion game is a game $G=\{N,R,\Sigma,P\}$ where $N$ is the
finite set agents, $R$ is the finite set resources, $P=\{P_r\}_{r
\in R}$ are the resource costs functions, where $P_r:[1, \ldots, n]
\to \mathbb{R}$ and $\Sigma=\times_{i \in N}\Sigma_i$, where
$\Sigma_i \subseteq 2^R$, is $i$'s strategy space. Agent $i$ selects
$s_i \in \Sigma_i$ and pays $\sum_{r \in s_i}P_r(c(s)_r)$, where
$c(s)_r = \sum_{j \in N}\mathbb{I}_{\{r \in s_j\}}$ is the number of
agents who select $r$ in $s$. In utility terms, the utility of agent
$i$ is $U_i(s) = -\sum_{r \in s_i}P_r(c(s)_r)$. If $\Sigma_i = R
\;\; \forall i \in N$ then $G=\{N,R,\Sigma,P\}$ is called a {\it
Simple Congestion Game}.
We will assume that the $P_r$ functions are non-negative and
increasing ($P_r(1)$ can be zero).
Fix a Simple Congestion Game or Coalitional Congestion Game (CCG) $G$ with $R$ resources and $n$ (sub-)agents. A
\emph{congestion vector} is an element of $\mathbb{N}^R$ whose
elements sum up to $n$. A Simple Congestion Game (CCG) $G$ and a strategy profile $s$
induce a congestion vector $c(s)$: $\{c(s)_r\}_{r \in R}$.
A strategy profile $s$ of a Coalitional Congestion Game $G^C$
induces a private congestion vector for each of the agents in $G^C$.
Such vector for agent $k$ will be $c_k$: $c_k(s_k)_r = |\{i \in C_k
: s_{k,i}=r\}|$ which is an element if $\mathbb{N}^R$ whose elements
sum up to $|C_k|$.
Let $X$ be a subset of the strategy profiles space. We denote
$c(X)$ as the corresponding set of congestion vectors: $\forall X
\subseteq S \;\; c(X) = \{c(s) \;s.t: s \in X \}$
\section{CCG and Pure Nash Equilibria}
The following preliminary result asserts that if the a Nash equilibrium the underlying game is composed of strategies such that all sub-agents of
any agent choose different resources, then it is also an equilibrium of the coalitional game.
\begin{proposition}\label{layer NE}
Let $G$ be a Simple Congestion Game , $C$ a partition of $N$ and $G^C$ be the induced CCG.
Let $s$ be a
strategy profile of $G^C$ where $s_{k,i} \neq s_{k,j} \;\; \forall
k \in N^C, \forall i,j \in C_k$. If $c(s) \in c(NE(G))$ then $s \in NE(G^C)$.
\end{proposition}
This is key to proving our central result about the existence of a Nash equilibrium in CCGs:
\begin{theorem}\label{Pairs Pure NE}
Let $G$ be a Simple Congestion Game , $C$ a partition where the largest element is of size 2, and $G^C$ a
CCG with the underlying game $G$ and the partition $C$. Then
$NE(G^C) \neq \emptyset$. That is, if the largest coalition is a Pair,
a Pure Nash equilibrium always exists.
\end{theorem}
Does this existence result extend to other partition forms, where the maximal element has
more than two sub-agents? The following example demonstrates that this is not true in general:
\begin{example}\label{No NE}
Consider a game with two identical resources A and B and four sub
agents, with the following payment functions:
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
Resource / Agents \#: & 1 & 2 & 3 & 4 \\
\hline
A: & 0 & 12 & 16 & 18 \\
\hline
B: & 0 & 12 & 16 & 18 \\
\hline
\end{tabular}
\end{center}
When $C=[\{1,2,3\},\{4\}]$ the matrix form of the resulting 2-player CCG is:
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$G^{C}$ & A & B\\
\hline
A,A,A & -54, -18 & -48, 0 \\
\hline
A,A,B & -32, -16 & -36, -12 \\
\hline
A,B,B & -36, -12 & -32, -16 \\
\hline
B,B,B & -48, 0 & -54, -18 \\
\hline
\end{tabular}
\end{center}
Whereas the underlying Simple Congestion Game has a pure Nash equilibrium this CCG has none. To verify this note for the compound agent (made up of 3 sub-agents) the strategies $AAA$ and $BBB$ are dominated. Following their deletion the remaining game is one of matching pennies and hence has no
Pure Nash equilibrium .
\end{example}
Can the result of Theorem \ref{Pairs Pure NE} be extended to CCGs
with small coalition size, but with an underlying congestion game
that is not simple? Again, the answer is negative:
\begin{example}\label{not SCG underlying}
Let $G$ be a Congestion Game with three identical resources and
three agents. Each agent of $G$ chooses two of the three
resources. The cost of each resource is $P(n)=6-\frac{6}{n}$.
Let $C = [\{1,2\}\{3\}]$, and $G^C$ is the CCG with underlying
game $G$ and partition $C$. After omitting identical strategies
due to sub agents symmetry $G^C$ looks as follows:
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
$G^C$ & AB & AC & BC\\
\hline
AB,AB & -16,-8 & -14,8 & -14,-4 \\
\hline
AC,AC & -14,-4 & -16,4 & -14,-4 \\
\hline
BC,BC & -14,-4 & -14,-4 & -16,-8 \\
\hline
AB,AC & -11,-7 & -11,-7 & -12,-6 \\
\hline
AB,BC & -11,-7 & -12,-6 & -11,-7 \\
\hline
AC,BC & -12,-6 & -11,-7 & -11,-7 \\
\hline
\end{tabular}
\end{center}
Note that compound agent's strategies (AB,AB), (AC,AC) and
(BC,BC) are dominated. Note that the remaining game has no pure Nash equilibrium.
\end{example}
\section{CCG and Potential}
An {\em Exact Potential} is a function $\mathbb{P}:S \to \mathbb{R}$ satisfying:
\begin{eqnarray}\label{Potential Definition}
\mathbb{P}(s) - \mathbb{P}(s_{-i}, t_i) =
U_i(s) - U_i(s_{-i},
t_i) \\ \nonumber \forall i \in N, \forall t_i \in S_i, \forall s \in S_1
\times S_2 \ldots \times S_n
\end{eqnarray}
Games with a potential function are called {\em Potential Games}. It is well known that potential games have a pure Nash equilibrium (see Monderer and Shapley \cite{Monderer Shapley 96}). In particular Congestion Games are potential games. Fotakis et al \cite{Fotakis} prove that a CCG,
where the cost functions of the resources of the underlying game are linear, is a potential game.
Removing the linearity assumption is problematic. In fact, even in
the case of a CCG with a maximal coalition of size $2$,
which guarantees the existence of a pure Nash equilibrium (Theorem
\ref{Pairs Pure NE}), the existence of Exact Potential is not
guaranteed. In the following example we show that the existence of a potential function implies linearity of the cost functions:
\begin{example}\label{no potential}
Consider a congestion games with 2 resources, $A$ and $B$, with costs $a_1, a_2, a_3$ and $b_1, b_2, b_3$, correspondingly. Assume there are 3 players and set the coalition structure to $C=[\{1,2\}\{3\}]$.
This induces the following two player CCG, given in matrix form:
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$G^{C}$ & A & B\\
\hline
A,A & $2a_3, a_3$ & $2a_2, b_1$ \\
\hline
A,B & $a_2+b_1, a_2$ & $a_1+b_2, b_2$\\
\hline
B,B & $2b_2, a_1$ & $2b_3, b_3$\\
\hline
\end{tabular}
\end{center}
Assume this game has an exact potential with the following values:
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$G^{C}$ & A & B\\
\hline
A,A & $P_1$ & $P_2$ \\
\hline
A,B & $P_3$ & $P_4$\\
\hline
B,B & $P_5$ & $P_6$\\
\hline
\end{tabular}
\end{center}
From the definition of exact potential the following must hold (see, in addition, Theorem 2.9. in Monderer and Shapley \cite{Monderer Shapley 96}):
\begin{eqnarray*}
(a_2+b_1-2a_3)+(a_3-b_1)+(2a_2-a_1-b_2)+(b_2-a_2)=\\ (P_3-P_1)+(P_1-P_2)+(P_2-P_4)+(P_4-P_3) = 0.\\
\end{eqnarray*}
Similarly:
\begin{eqnarray*}a_2+b_1-2b_2+a_1-b_3+2b_3-a_1-b_2+b_2-a_2=0\\
2a_3-2b_2+a_1-b_3+2b_3-2a_2+b_1-a_3=0.
\end{eqnarray*}
Manipulating these equalities leads to:
\begin{eqnarray*}
2a_2=a_1+a_3\\
2b_2=b_1+b_3,
\end{eqnarray*}
which implies that the cost functions are linear.
\end{example}
Using this example we can now prove our final result:
\begin{theorem}\label{linear costs}
Let $G$ be a Simple Congestion Game and let $C$ be a partition that has at least one element of
size 1 and at least one element of size 2. Let $G^C$ be a CCG with
the underlying game $G$ and partition $C$. $G^C$ will posses an
Exact Potential iff the CCG is linear.
\end{theorem}
\begin{thmproof}
Sufficiency - This has been obtained Fotakis et al. \cite{Fotakis} (Theorem 6).
Necessity - Recall that example \ref{no potential} provides a 3 player congestion game and a coalition structure that yields a CCG for which linear cost function are necessary for the existence of a potential.
The reason that the the linearity extends beyond the example to all situations implied in the theorem is that for any general CCG we can fix the strategy for all but 2 agents, of which one has 2 sub agents and one has a single sub agent. We can now look at the induced 2 player game. If the original game was a potential game so must be the induced game. The example then implies linearity in the induced game. However, as we can arbitrarily fix the strategy for all but the relevant 3 sub agents the result follows.
\end{thmproof}
\begin{appendix}
\section{Appendix - Omitted Proofs}
We begin with the definition of an auxiliary game. Let $G$ be a Simple Congestion Game and let $C$ be a partition. The {\em Restricted Coalitional Congestion Game}, denoted $\overline{G^C}$, is a CCG where coalitions are restricted strategies such that distinct sub-agents choose distinct resources. More formally:
\begin{definition}
Let $G$ be a Simple Congestion Game and let $C$ be a partition. The {\em Restricted Coalitional Congestion Game}, denoted $\overline{G^C}$, is the game $\overline{G^C}= \{N^c,\overline{S^c},U^c\}$, where $N^c$ and $U^c$
are as before and $\overline{S^c} = \{\overline{S^c_k}\}_{k \in C}$
where $\overline{S^c_k} = \{\times_{i \in C_k}S_{k,i} : s_{k,i}\neq
s_{k,j} \;\; \forall i,j \in C_k$\}.
\end{definition}
The following result about pure Nash equilibria in restricted coalitional congestion games will be useful for proving our main result:
\begin{lemma}\label{restricted layer NE}
Let $G$ be a Simple Congestion Game and $C$ a partition of $N$. Let $\overline{G^C}$ be
a Restricted CCG with the underlying game $G$ and the partition $C$.
Let $s$ be a strategy profile of $\overline{G^C}$ (where $s_{k,i}
\neq s_{k,j} \;\; \forall k \in N^C, \forall i,j \in C_k$). If $c(s)
\in c(NE(G)) \Rightarrow s \in NE(\overline{G^C})$
\end{lemma}
\begin{proof}
For two congestion vectors $u,v$, let $d(u,v) = \frac{\sum_{r\in R} |{v_r-u_r}|}{2}$, denote the distance between these two vectors.
Let $s$ be a strategy profile satisfying the condition of the lemma,
let $k$ be an arbitrary agent in the game $\overline{G^C}$ and
denote by $s_{-k}$ be the strategy profile of all players except
$k$. We denote by $BR(s_{-k})$ the set of $k$'s best reply
strategies to $s_{-k}$. Assume , by way of contradiction, that $s_k
\not \in BR(s_{-k})$ and let $t_k \in BR(s_{-k})$ be a best reply to
$s_{-k}$ which corresponding congestion vector has a minimal
distance to $c(s_k)$. Namely, $d(c(t_k),c(s_k)) \le
d(c(t'_k),c(s_k)) \ \ \forall t'_k \in BR(s_{-k})$.
In $\overline{G^C}$ each agent selects each resource at most once.
As $c_k(t_k) \neq c_k(s_k)$ this implies that there are two resources, say
$r$ and $x$, such that: $(c_k(t_k)_r, c_k(t_k)_x)=(1,0)$ and $(c_k(s_k)_r, c_k(s_k)_x)=(0,1)$ and, in addition, that $
c(s)_r = c(t_k,s_{-k})_r - 1$ and $c(s)_x = c(t_k,s_{-k})_x+1
$.
Let $i$ be the sub-agent of $k$ that chooses $r$ in $t_k$ but chooses a resource different than $r$ in $s_k$. Let $t'_k$ be a strategy for agent $k$, derived from $t_k$ by moving sub
agent $i$ from $r$ to $x$. This results in $c(t'_k,s_{-k})_x=c(s)_x$ and $c(t'_k,s_{-k})_r=c(s)_r$.
By assumption, $c(s) \in NE(G)$. Therefore $P_x(c(s)_x) \leq P_r(c(s)_r+1)$ and so:
\begin{equation}
P_x(c(t'_k,s_{-k})_x)=P_x(c(s)_x) \leq P_r(c(s)_r+1) =
P_r(c(t_k,s_{-k})_r)
\end{equation}
Thus, the contribution of sub agent $i$ to agent $k$'s payment in $(t'_k,s_{-k})$ is less or equal its contribution to $k$'s payment in $(t_k,s_{-k})$. As the only change in $k$'s strategy between $t_k$ and $t'_k$ is $i$'s choice, we conclude that $k$'s payment in $(t'_k,s_{-k})$ is less or equal its payment in $(t_k,s_{-k})$, and so $t'_k \in BR(s_{-k})$. However, by construction $d(c(t'_k),c(s_k)) \le d(c(t_k),c(s_k))$, contradicting the way $t_k$ was chosen.
\end{proof}.
{\bf Proof of Proposition \ref{layer NE}}:
Let $s$ be a profile as described in the Proposition. Let $t_k$ be the
best reply strategy for agent $k$ to $s_{-k}$. We show that
$c_k(t_k)_r \leq 1 \;\forall k \in N^C$ and $\forall r \in R$.
Assume this is not true and that some resource $r$,
$t_{k,i}=t_{k,j}=r$. Since $c_k(s_k)_r \leq 1$ we know that
$c(s_{-k},t_k)_r > c(s)_r$. Therefore there must exist some resource
$x$ such that $c(s_{-k},t_k)_x < c(s)_x$.
Let $t'_k$ be a strategy profile derived from $t_k$ by moving sub
agent $i$ from $r$ to $x$. In the strategy profile $(s_{-k},t'_k)$
agent $i$ pays $P_x(c(s_{-k},t'_k)_x)$. By construction:
\begin{equation}\label{31eq1}
c(s_{-k},t'_k)_x=c(s_{-k},t_k)_x+1 \leq c(s)_x
\end{equation}
From Equation \ref{31eq1} and the fact that $c(s) \in NE(G)$ we get
that:
\begin{equation}\label{31eq2}
P_x(c(s_{-k},t'_k)_x) = P_x(c(s_{-k},t_k)_x+1) \leq P_x(c(s_x)_x)
\leq P_r(c(s)_r+1)
\end{equation}
Using monotonicity of the cost functions and the fact that $c(s)_r+1
\leq c(s_{-k},t_k)_r$ we get:
\begin{equation}\label{31eq3}
P_r(c(s)_r+1) \leq P_r(c(s_{-k},t_k)_r)
\end{equation}
Combining Equations \ref{31eq2} and \ref{31eq3} we get that
\begin{equation}\label{31eq4}
P_x(c(s_{-k},t'_k)_x) \leq P_r(c(s_{-k},t_k)_r)
\end{equation}
We will now show that $k$ is better off in the strategy profile
$(s_{-k},t'_k)$ than in $(s_{-k},t_k)$, thus contradicting the fact
that $t_k$ is a best response to $s_{-k}$. We do this but analyzing
each of $k$'s sub-agents:
\begin{itemize}
\item
Sub agent $i$ pays in $(s_{-k},t'_k)$, where he chose $x$, no more
than than in $(s_{-k},t_k)$, where he chose $r$ (equation
\ref{31eq4}).
\item
From definition of $x$, $c_k(t_k)_x<c_k(s_k)_x$. Since
$c_k(s_k)_x=1$ and $c_k(s_k)_x>c_k(t_k)_x$, we get that
$c_k(t_k)_x=0$. Thus, apart from agent $i$ no other sub agent of $k$
chose $x$ in $t_k$.
\item
Sub agent $j$, who selects $r$ both in $(s_{-k},t'_k)$ and
$(s_{-k},t_k)$, pays strictly less in $(s_{-k},t'_k)$ than in
$(s_{-k},t_k)$, because $c(s_{-k},t'_k)_r<c(s_{-k},t_k)_r$. This
inequality holds for any other sub agent of $k$ who chose $r$ in
$t_k$
\item
All sub agents who choose a resource in the set $R \setminus\{r,x\}$
pay the same in $(s_{-k},t_k)$ and $(s_{-k},t'_k)$.
\end{itemize}
To conclude, agent $k$ pays strictly less in $(s_{-k},t'_k)$ than in
$(s_{-k},t_k)$. This contradicts the fact that $t_k$ is a best reply
to $s_{-k}$. Therefore, any best reply of $k$ to $s_{-k}$ must be
such that all the sub-agents choose different resources.
Thus, agent $k$'s best reply strategy to $s_{-k}$ is a strategy that
is allowed also in $\overline{G^C}$. We couple this observation with
the observation that $s$ is a Nash equilibrium of $\overline{G^C}$
(follows from Lemma \ref{restricted layer NE}) and the fact that $k$
is arbitrary to conclude that $s\in NE(G^C)$.
QED
{\bf Proof of Theorem \ref{Pairs Pure NE}:}
Let $s$ be an arbitrary Nash equilibrium of $G$
Case 1 - Assume that $c(s)_r \le |N^C|$ for all resources $r \in R$.
In this case we can re-arrange the players over the resources so
that the result will be a strategy profile with the same congestion
vector, and furthermore, for any $k \in N^C$ its two sub agents, $i,
j$, choose different resources. The resulting vector is also in
$NE(G)$ and complies with the conditions in Proposition \ref{layer NE}.
Therefore that proposition suggests that $s$ is a Nash equilibrium of
$G^C$.
Case 2 - Let us denote by $r$ the resource with the highest
congestion in $s$ and assume $c(s)_r>|N^C|$. We argue that without
loss of generality (by rearranging the players) $s$ has the
following two properties: (a) if agent $k$ has its 2 sub agents on
the same resource then it must be the case that the corresponding
resource is $r$, that is, $\forall k \in N^C$ $c(s_k)_{r'}>1$
implies $r'=r$; and (b) all agents have at least one sub-agent
choose $r$.
Note some properties of the strategy tuple $s$:
\begin{enumerate}
\item Let $k$ be an agent with a single sub agent. Then this sub agent must
be on $r$ and it has no profitable
deviation.
\item Let $k$ be an agent with a two sub agents, $i$ and $j$. Assume $i$
is on $r$ and $j$ is on some $r'\not = r$. Moving a single subagent cannot be
profitable.
\item Let $k$ be an agent with two sub-agents, $i$ and $j$. Assume $i$
is on $r$ and $j$ is on some $r'\not = r$. Moving both sub agents simultaneously cannot be profitable as at least one of these moves makes $k$ worse off, while the other cannot improve $k$
payoff.
\item Let $k$ be an agent with two sub-agents, $i$ and $j$, both on $r$.
Moving both sub agents cannot be
profitable.
\end{enumerate}
Thus, if $s$ is not a NE of $G^c$, the only profitable deviation possible is for an agent $k$ with two
sub-agents, $i$ and $j$, both on $r$, to move one sub-agent, say $j$ to another resource, say $r'$.
Furthermore, let us assume that this is the most profitable deviation for $k$. That is
$P_{r'}(c(s_{r'})+1) \leq P_{r''}(c(s_{r''})+1)$ for all $r'' \neq
r$. We denote the resulting strategy profile by $s'$. Note the
properties of $s'$:
\begin{enumerate}
\item All agents with a single sub-agent choose the resource $r$.
\item All agents with two sub agents, have at lease one sub agent in $r$.
\item The payment of all sub agents in $r$ is lower compared with $s$, while the payment of all subagents in $R\setminus \{r\}$ is at least as large compared with the payment in $s$.
\item Any agent with two sub agents, one on $r$ and one on some $r'' \not = r$ pays (weakly) less than what $k$ paid in $s$
\end{enumerate}
Assume $s'$ is not a Nash equilibrium, then there must be some profitable deviation. What are the possible profitable deviations?
\begin{enumerate}
\item Let $k'$ be an agent with a single sub agent. Then this sub agent must be on $r$ and it has no profitable deviation. Recall that the payment of $k$ is $s'$ is lower than the payment of $k'$ is $s$.
\item Let $k'$ be an agent with a two sub agents, $i$ and $j$. Assume $i$ is on $r$ and $j$ is on some $r'\not = r$. Moving $i$ cannot be profitable for the same argument as above. Moving $j$ to another resource in $R\setminus \{r\}$ cannot be profitable, so the only profitable deviation might be moving $j$ back to $r$. However, this would result $k'$ paying the same payment that $k$ paid in $s$, which by construction of $s'$ is higher than what $k$ pays in $s'$. Implying that $k'$ had a profitable deviation in $s$, thus contradicting what we already know.
\item Let $k$ be an agent with two sub-agents, $i$ and $j$. Assume $i$ is on $r$ and $j$ is on some $r'\not = r$. Moving both sub agents simultaneously cannot be profitable as at least one of these moves makes $k$ worse off, while the other cannot improve $k$ payoff.
\item Let $k$ be an agent with two sub-agents, $i$ and $j$, both on $r$. Moving both sub agents cannot be profitable.
\end{enumerate}
Once again, the only profitable deviation possible is for an agent $k'$ with
two sub-agents both on $r$, to move one sub-agent. The resulting
strategy profile $s''$, once more, only allows for profitable
deviations of the same form. Namely, for an agent $k'$ with two
sub-agents both on $r$, to move one sub-agent. We continue
iteratively in the same manner. As the process is bounded
by the number of agents selecting $r$ with both sub agents
in $s$, it must end in finitely many steps. The final strategy vector has no profitable deviations and is, therefore, a Nash equilibrium of the game.
QED
\end{appendix}
|
\section{Introduction}\label{sec:Introduction}
A symmetry group of a dynamical system may be discrete or continuous. The existence of discrete symmetries implies the existence of related sets of orbits and imposes constraints on bifurcations \cite{Chossat00, Golubitsky02}. Continuous symmetry often results in reduction; for example, the classical results of Sophus Lie are concerned with the reduction of order of an ODE or PDE with symmetry \cite{Olver93}. In the Hamiltonian or Lagrangian context, a continuous symmetry implies, through Noether's theorem, the existence of an invariant, and the reduction of the dynamics by two dimensions \cite{Marsden99,Marsden07,Holm09}.
In this paper, we are are interested in global reduction theory for maps that have continuous symmetries. A continuous symmetry of a map $f: M \to M$ is a vector field whose flow commutes with the map. The set of symmetries forms a Lie algebra. Symmetry reduction for maps seems to have been first studied by Maeda \cite{Maeda80, Maeda87}, who showed that a map with an $s$-dimensional, Abelian Lie algebra of symmetries can be written locally in a skew-product form
\beq{skew}
F(\sigma,\tau) = (k(\sigma), \tau+ \rot(\sigma)) \;,
\end{equation}
where $\sigma \in {\mathbb{ R}}^{n-s}$ and $\tau \in {\mathbb{ R}}^s$, in the neighborhood of any point where the rank of the symmetry group is $s$.
We will show in \Sec{mapSymmetry} that if the symmetry flow has a global Poincar\'{e} section $\Sigma$ that is relatively closed in a manifold $M$ (that is not necessarily compact), then we can find a covering map
of the form $p:\Sigma \times {\mathbb{ R}}\to M$. If some topological conditions are satisfied, then the map $f$ has a lift $F$ to $\Sigma \times {\mathbb{ R}}$ with the skew-product form \Eq{skew}, see \Th{Reduction}. The idea relies on the fact that the topologies of $\Sigma$ and $M$ may constitute an obstruction for this lift to exist. We call this procedure \emph{reduction by lifting}; it is a global version of Maeda's result.
This global reduction procedure is distinct from the general local reduction method due to Palais \cite{Palais61}. He showed that when an $s$-dimensional symmetry group has a proper action on $M$, then there is a codimension-$s$ local neighborhood of each point, a \emph{Palais slice}, such that the group orbit of a slice forms a tube within which the orbit trivializes: there are generalized ``flow-box" coordinates. In this paper, we are primarily concerned with one-parameter symmetry groups, though our results can be extended to the case of multi-dimensional, Abelian groups. Moreover, we will show that the skew-product \Eq{skew} can be globally valid on $M$, or at least on an open, dense subset.
In the classical theory of flows with Lie symmetry groups, a flow on a manifold $M$ with symmetry group $G$ that acts properly on $M$, can be reduced to a flow on the space of group orbits, $M/G$. When the action of $G$ is free, then the group orbit space is a manifold; alternatively $M/G$ is an ``orbifold" and the reduction has singularities.
By contrast, in our reduction procedure, there is no need to assume that the action of the symmetry group is proper or free; it thus circumvents some of the problems with singular reduction. A number of examples are given in \Sec{Examples}.
One motivation of our study is to extend some results from the setting of Hamiltonian vector fields or symplectic maps to the setting of divergence-free vector fields or volume-preserving maps. Indeed, whenever a dynamical system falls into a particular structural class (symplectic, volume preserving,...), the question arises whether the reduction by appropriate symmetries can be performed so that the structure is preserved.
We will show in \Sec{VP} that when reduction by lifting is applied to volume-preserving maps, then the reduced map $k$ of \Eq{skew} is also volume preserving with respect to a natural volume form on $\Sigma$. In particular, in the three-dimensional, volume-preserving case, the reduced map $k:\Sigma\to\Sigma$ is symplectic.
As we recall in \Sec{Invariant}, Noether's theorem implies that whenever a Hamiltonian flow has a Hamiltonian symmetry there is an invariant, and conversely, that every invariant generates a Hamiltonian vector field that is a symmetry. This result also holds for symplectic twist maps with a Lagrangian generating function \cite{Logan73, Maeda81, Wendlandt97, Mansfield06}.
We will generalize a result of Bazzani \cite{Bazzani88} to show that, with the addition of a recurrence condition, Noether's theorem also applies more generally to symplectic maps. However, it is important to note that symmetries do not generally lead to invariants nor do invariants necessarily give rise to symmetries.
In \Sec{Comparison} we compare our procedure with two standard reduction procedures, using as an example a four-dimensional symplectic map with rotational symmetry. We discuss the advantages and shortcomings of reduction by lifting.
\section{Symmetry Reduction by Lifting}\label{sec:mapSymmetry}
In this section we investigate the conditions under which a map or flow that has a symmetry can be written in the skew-product form \Eq{skew}. Whenever this is possible,
the process reduces the effective dimension of the dynamics by one. To accomplish this reduction, we suppose that the symmetry vector field generates a complete flow that admits a global Poincar\'{e} section. We recall below the relationship between the existence of cross sections, covering spaces and fundamental groups and show how to use it to obtain the symmetry reduction.
\subsection{Invariants and Symmetries}\label{sec:notation}
We start by recalling some basic notation (see also \App{notation}) and facts about symmetries and invariants of vector fields and maps.
We suppose that $M$ is a (not necessarily compact) $n$-dimensional, path-connected manifold and denote the set of $C^1$ vector fields on $M$ by $\mathcal{V}(M)$ and the set of $k$-forms on $M$ by ${\Lambda}^k(M)$. The flow, $\varphi_{t}(x)$, of a vector field $X \in \mathcal{V}(M)$ is the solution of the initial value problem
$\dot \varphi_{t}(x) = X(\varphi_{t}(x))$ with $\varphi_{0}(x)=x$. In this paper we will assume that each vector field has a complete flow; that is, $\varphi_t: M \to M $ is a diffeomorphism for all $t \in {\mathbb{ R}}$. In this case, $\varphi$ is an action of the one-dimensional Lie group ${\mathbb{ R}}$ on $M$ with composition rule $\varphi_t\circ \varphi_s=\varphi_{s+t}$.
An \emph{invariant} for a vector field $X$ is a function $I: M \to {\mathbb{ R}}$ such that
$
L_X(I) = i_{X}dI = 0 \;,
$
where $L_X$ is the Lie derivative (see \App{notation}). Equivalently $I$ is invariant when $\varphi^*_t I = I$.
Similarly $I$ is an invariant for a map $f: M \to M$ if
\beq{MapInvariant}
f^* I := I\circ f = I \;.
\end{equation}
A \emph{continuous symmetry} of a vector field $X$ is a vector field $Y$ that commutes with $X$, $[Y,X] = 0$, or equivalently, $L_Y X = 0$.
This implies that the flows of $X$ and $Y$, say $\varphi$ and $\psi$ respectively, commute \cite{Arnold78}:
\[
\varphi_t \circ \psi_s = \psi_s \circ \varphi_t \;.
\]
Similarly, a vector field $Y$ is a symmetry of a map $f$ if
\beq{MapSymmetry}
f^*Y = Y \;,
\end{equation}
where $f^*Y := ( Df(x) )^{-1} Y(f(x))$.
This means that the transformation $f$ leaves the differential equation $\dot y = Y(y)$ invariant. Since $y$ and $z=f(y)$ satisfy the same system of differential equations, the symmetry extends to the Lie group generated by the flow of $Y$:
\[
\psi_t \circ f = f \circ \psi_t \;.
\]
Thus $f$ is an equivariant transformation of the flow.
Clearly, the collection of vector fields that are symmetries of a map form a Lie algebra. For example if $X,Y\in\mathcal{V}(M)$ are symmetries of $f$, then $f^*[X,Y]=[f^*X,f^*Y]=[X,Y]$, so $[X,Y]$ is also a symmetry.
\subsection{Global Poincar\'{e} Sections}
We will assume $f:M \to M$ is a diffeomorphism with a symmetry vector field $Y$, \Eq{MapSymmetry}. Standing assumptions are that $M$ is an $n$-dimensional, path-connected manifold, and that $Y$ has a complete flow with a global Poincar\'{e} section.
Recall that a relatively closed, codimension-one submanifold $\Sigma \hookrightarrow M$
is a \emph{Poincar\'{e} section} or cross section of a flow $\psi$ if it is transverse to the flow, and is a \emph{global Poincar\'{e} section} of a complete flow if every orbit of the flow has both forward and backward transversal intersections with $\Sigma$, see the left pane of \Fig{cross}. Recall that $\Sigma$ is relatively closed in $M$ if, given a sequence in $\Sigma$ that converges in $M$, the sequence also
converges in $\Sigma$.
\InsertFig{cross.pdf}{A global Poincar\'{e} section $\Sigma$ has two representations that are equivalent by \Th{covering}:
\emph{a}) as an embedding in the original manifold $M$, and
\emph{b}) lifted to a cover $C = \Sigma \times {\mathbb{ R}}$.
In both cases, an orbit of the flow $\psi$ is drawn.}{cross}{3.5in}
When $\Sigma$ is a global Poincar\'{e} section for $\psi$, the first return time to $\Sigma$, $T: \Sigma \to {\mathbb{ R}}^+$, is the smallest positive number such that for, each $\sigma \in \Sigma$, $\psi_{T(\sigma)}(\sigma)\in\Sigma$. In fact, $T$ is continuous.
The first return map $r_\psi: \Sigma \to \Sigma$, also known as the Poincar\'{e} map, is defined by
\beq{PoincareMap}
r_\psi(\sigma) := \psi_{T(\sigma)}(\sigma) \;;
\end{equation}
it is a diffeomorphism. The sequence of return times $T_n$, $n \in {\mathbb{ Z}}$, is defined so that $T_0(\sigma)=0$,
\[
T_{n+1}(\sigma):=T_n(\sigma)+T(r_\psi^n(\sigma)) \;, \quad n \ge 0 \;,
\]
and $T_{-n}(\sigma)=-T_n(r_\psi^{-n}(\sigma))$.
Using this notation, the iterates of the first return map are
\beq{PoincareIterates}
r_\psi^n(\sigma)=\psi_{T_n(\sigma)}(\sigma) \;,
\end{equation}
for all $n \in {\mathbb{ Z}}$. Note that the sequence of return times is strictly increasing with $n$, and must be unbounded when the section $\Sigma$ is relatively closed in $M$.
Indeed, if for some $\sigma_0$ the sequence $T_n(\sigma_0)$ were bounded then, since it is increasing, it would converge. Consequently, the difference $T_{n+1}(\sigma_0)-T_n(\sigma_0) \to 0$ and therefore by definition
\[
T(r_\psi^n(\sigma_0))\to 0
\]
as $n\to\infty$. Thus $r_\psi^n(\sigma_0)=\psi_{T_n(\sigma_0)}(\sigma_0)\to\sigma^*$ would also converge, and since $\Sigma$ is relatively closed, we could conclude that $\sigma^*\in\Sigma$. By continuity, $T(\sigma^*)=0$, but this contradicts the fact that the return times are always positive. A similar argument shows that $T_n(\sigma)$ is unbounded when $n\to -\infty$.
A global Poincar\'{e} section $\Sigma$ can be viewed both as a submanifold of $M$, and---as illustrated in the right pane of \Fig{cross}---as the base of a covering space for $M$.
Recall that a manifold $C$ is a cover of $M$ if there is a differentiable, surjective function $p:C\to M$ such that each $m\in M$ has a neighborhood $\mathcal{U}$ for which $p^{-1}(\mathcal{U})$ is the disjoint union of a countable number of open sets in $C$, each of which is diffeomorphic, via $p$, to $\mathcal{U}$. The covering map $p$ is necessarily a local diffeomorphism. A covering space is a fiber bundle with a discrete fiber.
We will often think of $\Sigma$ both as a submanifold of $M$ and a space in its own right. Technically, we define an inclusion map $\iota:\Sigma\hookrightarrow M$ to express the embedding, but when there is little risk of confusion, we will let $\Sigma \subset M$ denote $\iota(\Sigma)$.
A fundamental tool for the rest of the paper is the following theorem of Schwartzman that shows how global Poincar\'{e} sections are related to covering spaces.
\begin{thm}[\cite{Schwartzman62}]\label{thm:covering}
A relatively closed submanifold $\Sigma$ is a global Poincar\'{e} section of a flow $\psi$ on a manifold $M$
if and only if the map $p:\Sigma\times{\mathbb{ R}}\to M$ defined by
\beq{covering}
p(\sigma,\tau)=\psi_\tau(\sigma)
\end{equation}
is a smooth cover of $M$ with an infinite, cyclic group of deck transformations.
\end{thm}
Recall that a deck transformation of $p$ is a map $\Delta:C\to C$ on the covering space such that $p \circ \Delta = p$; that is, deck transformations are the lifts of the identity \cite{Dieck08}.
For the global Poincar\'{e} section $\Sigma$, is easy to verify that
\begin{equation}\label{eq:Delta}
\Delta(\sigma,\tau)=(r_\psi(\sigma), \tau-T(\sigma)) \;,
\end{equation}
is a deck transformation on $\Sigma \times {\mathbb{ R}}$. Taking into account \Eq{PoincareIterates}, the iterates of $\Delta$ satisfy
\begin{equation}\label{eq:deck}
\Delta^n(\sigma,\tau)=(r_\psi^n(\sigma), \tau-T_n(\sigma))
\end{equation}
and are also deck transformations. Hence, the collection
$\{ \Delta^n: n\in {\mathbb{ Z}} \}$, is the cyclic group mentioned in \Th{covering}.
We next note that the orbits of $\Delta$ correspond to the fibers of the cover.
\begin{lem}\label{lem:Fiber}
For each $x \in M$, the fiber, $p^{-1}(x)$, of the covering map $p$ is an orbit of $\Delta$. In other words, if $p(\sigma_0,\tau_0) = x$, then
\[
p^{-1}(x)={\rm orb}_\Delta(\sigma_0,\tau_0)
:= \{\Delta^n(\sigma_0,\tau_0): n\in{\mathbb{ Z}}\} \;.
\]
\end{lem}
\begin{proof}
For any point $(\sigma',\tau') \in {\rm orb}_\Delta(\sigma_0,\tau_0)$, there is an $n$ such that $(\sigma',\tau') = \Delta^n(\sigma_0,\tau_0)$, and since $\Delta^n$ is a deck transformation, $p(\sigma',\tau') = p \circ \Delta^n(\sigma_0,\tau_0) = p(\sigma_0,\tau_0) = x$.
Conversely, suppose that $(\sigma^\prime,\tau^\prime)\in p^{-1}(x)$, i.e.,
\(
\psi_{\tau^\prime}(\sigma^\prime)=x=\psi_{\tau_0}(\sigma_0),
\)
so that $\sigma'= \psi_{\tau_0-\tau'}(\sigma_0) \in \Sigma$.
Assuming, without loss of generality, that $\tau_0>\tau^\prime$, then since the sequence $T_n(\sigma_0)$ is strictly increasing and unbounded, there must be an $n\in {\mathbb{ N}}$ so that
\[
T_n(\sigma_0)\leq \tau_0-\tau'<T_{n+1}(\sigma_0) \;.
\]
If we let $\sigma_n=r_\psi^n(\sigma_0)$ and $\kappa=\tau_0-\tau' -T_n(\sigma_0)$,
then $0\leq \kappa<T(\sigma_n)$ and $\sigma'=\psi_\kappa(\sigma_n)$. By definition of
the first return time, the only possibility is that $\kappa=0$.
Therefore, $\tau_0-\tau' = T_n(\sigma_0)$. Consequently, $\sigma' = r^n_\psi(\sigma_0)$
and thus $(\sigma',\tau') = \Delta^n(\sigma_0,\tau_0)$.
\end{proof}
\subsection{Symmetry Reduction by Lifting}
We now ask whether it is possible to lift $f$ to a cover $C$ in which the map takes the skew-product form \Eq{skew}. Recall that a lift of $f$ is a map
$F:C\to C$ such that
\beq{liftDef}
f\circ p=p\circ F \;.
\end{equation}
In particular, any deck transformation is the lift of the identity on the base.
Lifts are not unique, but all lifts are the same up to deck-transformations:
if $G$ is another lift of $f$ then $ G = \Delta^m F$ for some integer $m$.
Suppose that $x_0,y_0\in C$ are two points in the cover such that $f(p(x_0))=p(y_0)$.
A necessary and sufficient condition \cite{Massey} for the existence of a lift is that
\begin{equation}\label{eq:lift}
f_*p_*{\pi_1}\left(C,x_0\right)\subseteq p_*{\pi_1}\left(C,y_0\right) \;,
\end{equation}
where ${\pi_1}(C,x_0)$ is the fundamental group of $C$ based at $x_0$.
In addition, if the lift exists, it can be taken to satisfy $F(x_0)=y_0$.\footnote
{Recall that the fundamental group of a topological space $M$ depends on the choice of a base point $\xi_0$. However, the choice of base point is irrelevant provided the space $M$ is path-connected.}
In other words, each loop in $M$ that is a projection of a noncontractable loop in $C$ must map, under $f$, to another loop in the same projection. This is a nontrivial requirement because $p_*({\pi_1}(C,x_0))$ is necessarily a subgroup of ${\pi_1}(M,x_0)$. Note, however that the requirement would be trivially satisfied if the
cover were simply connected, since in that case ${\pi_1}(C,x_0)$ would be trivial.
\InsertFig{lift}{Failure of \Eq{lift} in an attempt to construct a lift of the cat map to the cylinder.}{lift}{3.8in}
The necessity of the condition \Eq{lift} is easy to illustrate. Consider the example of Arnold's cat map on the two-torus $M={\mathbb{ R}}^2/{\mathbb{ Z}}^2$: $f(x,y)=(2x+y,x+y)$. Does $f$ have a lift to the cylinder $C={\mathbb{ R}}\times{\mathbb{ S}}^1$ with the natural covering map, $p(z,\tau) = (z \mod 1, \tau)$?
The fundamental group ${\pi_1}(M,0)$ has two generators, say $\gamma_1,\gamma_2$; however, ${\pi_1}(C,0)$ has only one generator, say $\delta$. Since $p_*(\delta)=\gamma_2$, and $f_*(\gamma_2)=\gamma_1+\gamma_2$, the condition
\Eq{lift} is not satisfied. The geometric reason this fails is illustrated in \Fig{lift}.
Any loop $f\circ p\circ\delta$ has the homotopy type of $\gamma_1+\gamma_2$. However,
any lift of $f\circ p\circ\delta$ can not be a loop on $C$.
We will now give conditions for existence of a lift of $f$ on the cover $C=\Sigma\times{\mathbb{ R}}$ with $p:C\to M$ as in \Eq{covering}. Let $\Psi$ denote the trivial flow on $\Sigma\times{\mathbb{ R}}$,
\beq{trivial}
\Psi_t(\sigma,\tau)=(\sigma,\tau+t) \;.
\end{equation}
Clearly, $\Psi$ commutes with $\Delta$ and
\(
p\circ \Psi_t(\sigma,\tau)=p(\sigma,\tau+t)
=\psi_{\tau+t}(\sigma)=\psi_t\circ p(\sigma,\tau)
\)
so that $\Psi$ is a lift of $\psi$, in the sense that each $\Psi_t$ is a lift of $\psi_t$, for all $t$.
\begin{thm}[Symmetry Reduction]\label{thm:Reduction}
Suppose that a diffeomorphism $f:M\to M$ has a symmetry $\psi_t$ with global Poincar\'{e} section $\Sigma$
that is relatively closed in $M$ and let $p:\Sigma\times {\mathbb{ R}}\to M$ be the induced cover \Eq{covering}.
Let $\sigma_0\in\Sigma$ and $(\bar{\sigma},\bar{\tau})\in p^{-1}(f(\sigma_0))$.
Then the following statements are equivalent.
\begin{enumerate}
\item\label{part1} %
There exists a lift $F:\Sigma \times {\mathbb{ R}}\to \Sigma \times {\mathbb{ R}}$ of $f$ to the cover of the form
\beq{skewProduct}
F: \left\{ \begin{array}{l}
\sigma' = k(\sigma) \;, \\
\tau' = \tau + \rot(\sigma) \;,
\end{array} \right.
\end{equation}
where $k$ is a diffeomorphism of $\Sigma$, and $k(\sigma_0)=\bar{\sigma}$, $\rot(\sigma_0)=\bar{\tau}$.
\item\label{part2} If $\iota:\Sigma\hookrightarrow M$ is the standard inclusion, then
\beq{groupSubset}
f_*\iota_*{\pi_1}\left(\Sigma,\sigma_0\right)\subseteq
p_*{\pi_1}\left(\Sigma\times{\mathbb{ R}},(\bar{\sigma},\bar{\tau})\right) \;.
\end{equation}
\end{enumerate}
\end{thm}
\begin{proof}
\ref{part1})$\Longrightarrow$\ref{part2})
By \Eq{skewProduct}, $F(\sigma,0)=(k(\sigma),\rot(\sigma))$, and since $F$ is a lift of $f$, $f(\sigma)=f\circ p(\sigma,0)=p(k(\sigma),\rot(\sigma))$.
Let $G:\Sigma\to\Sigma\times{\mathbb{ R}}$ be given by $G(\sigma):=(k(\sigma),\rot(\sigma))$.
Then $p\circ G=f\circ \iota$ and $G(\sigma_0)=(\bar{\sigma},\bar{\tau})$. Consequently $f_*\iota_*{\pi_1}\left(\Sigma,\sigma_0\right)=p_*G_*{\pi_1}\left(\Sigma,\sigma_0\right)$, and since $G_* {\pi_1}(\Sigma,\sigma_0) \subseteq {\pi_1}\left(\Sigma\times{\mathbb{ R}},(\bar{\sigma},\bar{\tau})\right)$, and $p_*$ is injective, this directly implies \ref{part2}).
\ref{part2})$\Longrightarrow$\ref{part1})
Under the condition \Eq{groupSubset}, standard theorems of algebraic topology \cite{Massey} imply that there exists a function $G:\Sigma\to\Sigma\times{\mathbb{ R}}$ such that $p\circ G=f\circ \iota$. Define $k$ and $\omega$ by $G(\sigma)=(k(\sigma),\rot(\sigma))$. Since $f\circ \iota(\sigma_0)=p(\bar{\sigma},\bar{\tau})=f(\sigma_0)$, we can take $G(\sigma_0) = (k(\sigma_0), \omega(\sigma_0)) = (\bar{\sigma},\bar{\tau})$.
Now, for each $\tau\in{\mathbb{ R}}$, we define $F(\sigma,\tau):=\Psi_\tau(G(\sigma))$, where $\Psi$ is the trivial
flow \Eq{trivial}. We claim that $F$ is a lift of $f$, i.e., it satisfies \Eq{liftDef}. To show this, we notice that
\[
p\circ F(\sigma,\tau)=p\circ\Psi_\tau(G(\sigma))=\psi_\tau\circ p(G(\sigma))=
\psi_\tau(f(\sigma))=f(\psi_\tau(\sigma))=f\circ p(\sigma,\tau).
\]
\end{proof}
\InsertFig{SkewProduct}{Showing the ideas behind the skew-product form \Eq{skewProduct}.}{SkewProduct}{3.5in}
The geometry underlying the reduction \Eq{skewProduct} is illustrated in \Fig{SkewProduct}. Theorem \ref{thm:Reduction} can be seen to be equivalent to a commuting diagram.
Let $j:\Sigma\to \Sigma \times {\mathbb{ R}}$ be the inclusion given by $j(\sigma):=(\sigma,0)$.
When $\iota:\Sigma\hookrightarrow M$ is the standard inclusion,
then $p\circ j=\iota$.
Let $\Pi:\Sigma\times {\mathbb{ R}}\to \Sigma$ be the canonical projection.
Clearly, $\Pi\circ j=id_\Sigma$.
With this notation, \Th{Reduction} can be represented by the commutative diagram
\bsplit{commutative}
\xymatrixcolsep{4pc}
\xymatrix{
&\Sigma\ar[r]^k&\Sigma\\
\Sigma \ar[r]^j\ar@{_{(}->}[dr]_\iota\ar[ur]^{id_\Sigma}&
\Sigma\times{\mathbb{ R}} \ar[u]_\Pi\ar [r]^F\ar[d]_p&
\Sigma\times{\mathbb{ R}} \ar[u]_\Pi\ar[d]_p\\
&M\ar[r]^f&M}
\end{split}\end{equation}
In particular, the reduced map $k$ on $\Sigma$ can be written as
\beq{reducedMap}
k=\Pi\circ F\circ j \;.
\end{equation}
Furthermore, the choice of $\sigma_0$ and $(\bar{\sigma},\bar{\tau})$
such that $f(\sigma_0)=\psi_{\bar{\tau}}(\bar{\sigma})$, fixes the choice of
$k$ and $\rot$.
A simplification in \Th{Reduction} occurs if $f$ has a fixed point at a point $\sigma_0 \in \Sigma$. Since the inclusion $j$ induces an isomorphism of the fundamental groups,
$j_*:{\pi_1}\left(\Sigma,\sigma_0\right)\to{\pi_1}\left(\Sigma\times{\mathbb{ R}},(\sigma_0,0)\right)$,
\[
\iota_*{\pi_1}\left(\Sigma,\sigma_0\right)
=p_*j_*{\pi_1}\left(\Sigma,\sigma_0\right)
=p_*\left({\pi_1}\left(\Sigma\times{\mathbb{ R}},(\sigma_0,0)\right)\right) \;.
\]
Upon taking $\bar{\tau}=0$ and $\bar{\sigma}=\sigma_0$, the condition \Eq{groupSubset} then reduces to
\beq{newSubset}
f_*\iota_*{\pi_1}\left(\Sigma,\sigma_0\right) \subseteq
\iota_*{\pi_1}\left(\Sigma,\sigma_0\right)\;.
\end{equation}
Finally, we note a convenient equation determining $k$ and $\omega$ is obtained by combining the definition of the lift \Eq{liftDef}, the form of the covering map \Eq{covering}, and the skew-product form \Eq{skewProduct} to obtain
$
f( \psi_\tau( \sigma) ) = \psi_{\tau + \rot(\sigma)} ( k(\sigma) ) \;.
$
Since $f$ commutes with $\psi_\tau$ we find
\beq{kDefine}
k(\sigma) = \psi_{-\rot(\sigma)} ( f(\sigma) ) \in \Sigma \;.
\end{equation}
This determines $k$ and $\rot$ (up to the choice of lift) by the requirement that the right side must be in $\Sigma$.
\subsection{Deck symmetries and a homotopy invariant}
In this section we will show that lifts of maps with symmetries have a homotopy invariant.
\begin{lem}\label{lem:homotopy}
Suppose that $f$ and its lift $F$ satisfy the hypotheses of \Th{Reduction}, and that $\Delta$ is the deck transformation defined in \Eq{Delta}. Then there exists $m\in {\mathbb{ Z}}$ such that
$F\circ \Delta=\Delta^m\circ F$ and
the following identities are satisfied
\begin{align*}
k\circ r_\psi&=r_\psi^m\circ k \;,\\
\rot(\sigma)-\rot(r_\psi(\sigma))&=T_m(k(\sigma))-T(\sigma) \;.
\end{align*}
The integer $m$ is a homotopy invariant of the map $f$.
In addition, if $F$ is homotopic to the identity then $m=1$.
\end{lem}
\begin{proof}
Consider the map $G=F\circ \Delta\circ F^{-1}$. Clearly $p\circ G=p$, so $G$ has to be a deck transformation. This implies that there exists $m\in {\mathbb{ Z}}$ such that $G=\Delta^m$.
The integer $m$ is independent of the choice of lift. Indeed, suppose that $\tilde F$ is another lift of $f$, then since $p \circ F = p \circ \tilde F$, it follows from \Lem{Fiber} that there is an integer $j$ such that for any $(\sigma,\tau)$, $\tilde F(\sigma,\tau) = \Delta^j \circ F(\sigma,\tau)$. By continuity, $j$ is independent of $(\sigma,\tau)$, and consequently any two lifts differ at most by a deck transformation. Thus $\tilde F \circ \Delta = \Delta^{j+m} F = \Delta^m \tilde F$.
If $F$ is homotopic to the identity then $\Delta^{-1}\circ G$ is a deck transformation that is homotopic to the identity. The only possibility is that $\Delta^{-1}\circ G$ is the identity and therefore $m=1$. Similarly since the integer $m$ is independent of the lift, continuity implies any two homotopic maps will have the same value of $m$.
For the rest, it is enough to use \Eq{deck} in the skew product \Eq{skewProduct}.
\end{proof}
\subsection{Reduction of flows}\label{sec:FlowReduction}
Theorem \ref{thm:Reduction} also applies to a flow $\varphi$, with a symmetry $\psi$.
As in the theorem, we assume that $\psi$ has a global Poincar\'{e} section $\Sigma$.
\begin{cor}[Symmetry Reduction of Flows]\label{cor:FlowReduction}
Suppose that $\psi,\varphi$ are a pair of commuting flows on $M$ and that $\psi$ has a global Poincar\'{e} section $\Sigma$ that is relatively closed in $M$. Then there is a lift $\Phi_t$ of $\varphi_t$ to $\Sigma \times {\mathbb{ R}}$ of the form
\beq{FlowReduction}
\Phi_t(\sigma,\tau)=(k_t(\sigma), \tau+\rot(\sigma,t)) \;.
\end{equation}
In addition, $\Phi_t$ can be chosen to be a flow on $\Sigma\times{\mathbb{ R}}$ and hence the functions $k_t$ and $\rot(\cdot,t)$ satisfy
\begin{align*}
k_{t+s}&=k_t\circ k_s,\\
\rot(\sigma,t+s)&=\rot(\sigma,t)+\rot(k_t(\sigma),s) \;.
\end{align*}
\end{cor}
\begin{proof}
Just as in the definition \Eq{covering} of the cover $p$, let $q:\Sigma\times{\mathbb{ R}}\to M$ denote the map $q(\sigma,t)=\varphi_t(\sigma)$. We notice that
$q\circ j=p\circ j=\iota$. Hence for each $\sigma_0\in\Sigma$,
\[
q_*{\pi_1}\left(\Sigma\times{\mathbb{ R}},(\sigma_0,0)\right)=
q_*j_*{\pi_1}\left(\Sigma,\sigma_0\right)=
p_*j_*{\pi_1}\left(\Sigma,\sigma_0\right)=
p_*{\pi_1}\left(\Sigma\times{\mathbb{ R}},(\sigma_0,0)\right) \;.
\]
Consequently, there exists a map $Q:\Sigma\times{\mathbb{ R}}\to\Sigma\times{\mathbb{ R}}$ such that
$q= p \circ Q$. Equivalently, there exist functions $k(\sigma,t)$ and $\rot(\sigma,t)$ such that $Q(\sigma,t)=(k(\sigma,t),\rot(\sigma,t))$ and
$\varphi_t(\sigma)=\psi_{\rot(\sigma,t)}(k(\sigma,t))$. Denoting $k_t:=k(\cdot,t)$, define $\Phi_t$ by \Eq{FlowReduction}.
For each $t$, $\Phi_t$ is clearly a lift of $\varphi_t$, since
\[
p\left(\Phi_t(\sigma,\tau)\right)=\psi_{\tau+\rot(\sigma,t)}(k(\sigma,t))=
\psi_{\tau}\varphi_t(\sigma)=
\varphi_t\psi_{\tau}(\sigma)=
\varphi_t\left(p(\sigma,\tau)\right).
\]
We must show that it is possible to choose $\Phi_t$ so that $\Phi_0 = id_{\Sigma \times {\mathbb{ R}}}$,
and $\Phi_{t+s} = \Phi_t \circ \Phi_s$. Note first that $\Phi_0$ is a lift of the identity map
and so it must be a deck transformation. If $\Phi_0$ is not itself the identity, then we can
replace $\Phi_t$ by $\Phi_t \circ \Phi_0^{-1}$. By \Lem{homotopy}, since $\Phi_t$ is a map
that is homotopic to the identity, it commutes with deck transformations; thus the new $\Phi_t$
is still a lift of $\varphi_t$, and satisfies $\Phi_0 = id_{\Sigma \times {\mathbb{ R}}}$.
Now, for any $s \in {\mathbb{ R}}$, let $G_t = \Phi_{t+s} \circ \Phi_t^{-1} \circ \Phi_s^{-1}$.
Note that $p \circ G_t = p$, so that $G_t$ is a deck transformation. Moreover, since
$G_0 = id_{\Sigma \times {\mathbb{ R}}}$, $G_t$ is homotopic to the identity, and thus must be the
identity. In this way we conclude that $\Phi_t$ satisfies the group property, and is a flow.
\end{proof}
The implication of \Cor{FlowReduction} is that $k_t$ is a flow on the section $\Sigma$.
\subsection{Symmetry with an Invariant}\label{sec:InvariantSymmetry}
Recall that generally the existence of a symmetry does not imply that of an associated invariant. However, it is possible that both the map and the symmetry do share an invariant. In that case, we can see that the reduced map has the same invariant.
\begin{lem}\label{lem:Invariant}
Suppose $f:M\to M$ is a map with a symmetry $\psi$ and the hypotheses
of \Th{Reduction} hold. If $I:M\to {\mathbb{ R}}$ is an invariant of $f$ that is also an invariant of $\psi$, then $\iota^*I$ is an invariant of $k$.
\end{lem}
\begin{proof}
Notice that
\(
F^*p^*I=p^*f^*I=p^*I.
\)
Moreover, for any $(\sigma,\tau)\in\Sigma\times{\mathbb{ R}}$, \Eq{covering} implies
\[
p^*I(\sigma,\tau)=I(\psi_\tau(\sigma))=I(\iota(\sigma))=
I(\iota(\Pi(\sigma,\tau)))=\Pi^*\iota^*I(\sigma,\tau) \;.
\]
Using the expression \Eq{reducedMap} we can then conclude that
\[
k^*\iota^*I = j^*F^*\Pi^*\iota^*I
= j^*F^*p^*I = j^*p^*I
= \iota^*I \;.
\]
\end{proof}
Examples that have both invariants and symmetries will be given in \Sec{Examples}.
\subsection{Circle actions}\label{sec:CircleAction}
We have assumed that the symmetry vector field $Y$ generates a flow $\psi_t$ that corresponds to an action of the Lie group ${\mathbb{ R}}$ on the manifold $M$. However, in many cases, this action is periodic, and can thus be thought of as an action of the group ${\mathbb{ S}}^1$ on $M$. Typically, the temporal period of the orbit of $\psi_t$ will depend upon the point. Nevertheless, if $\psi_t$ has a global section $\Sigma$, and every orbit is periodic, then some iterate of the Poincar\'{e} map, $r_\psi$, \Eq{PoincareMap}, must be the identity, because the flow returns to the original point on the same circle. In this case there is a smallest positive integer $\ell$ such that $r_\psi^\ell= id_{\Sigma}$, for all $\sigma \in \Sigma$.
The existence of a global section for a circle action $\psi_t$ on $M$ is a
strong restriction on the topology of the manifold. Indeed, since the $\ell$-th iterate of the Poincar\'e map is the identity, a ${\mathbb{ Z}}_\ell \equiv {\mathbb{ Z}} \setminus \ell {\mathbb{ Z}}$ covering of $M$ is a trivial bundle over ${\mathbb{ S}}^1$, i.e.,\ $M = ( \Sigma \times {\mathbb{ S}}^1) / {\mathbb{ Z}}_\ell$.
In general the existence of a (free and proper) circle action gives $M$ the structure of a principal ${\mathbb{ S}}^1$-fiber bundle, though it need not be trivial. A classical example is provided by the Hopf fibration of ${\mathbb{ S}}^3$. At first it seems as if the reduction of \Th{Reduction} could not be done for a symmetry that is a non-trivial bundle that is not a discrete quotient of a direct product with ${\mathbb{ S}}^1$. However, one can often modify $M$ in a way that it acquires the necessary topology. This may be achieved by removing parts of $M$ that are invariant under the symmetry flow and under the map; examples are given \Sec{Examples} and \Sec{Comparison}. Some global topology may be lost because the modified $M$ will in general not be compact.
The question can also be inverted: how can one possibly achieve reduction {\em and}
reconstruction in the case where the symmetry group action induces a non-trivial fiber
bundle? The semi-direct product structure of \Eq{skewProduct} requires that the dynamics in the fiber are driven by the dynamics in the base. However, if it is not possible to globally define an origin in the ${\mathbb{ S}}^1$ fiber, this cannot be done.
\subsection{Orbit spaces and global Poincar\'{e} sections}\label{sec:OrbitSpaces}
Classical symmetry reduction begins with a smooth manifold $M$ (say without boundary, or even better a compact manifold) and a Lie group $G$ with a smooth and proper $G$-action $\Phi : G \times M \to M$. Two points on the same group orbit are now considered equivalent, and symmetry reduction means to study dynamics on the \emph{orbit space} $M/G$. In general this quotient is not a manifold, but just an orbifold. Each point $x \in M$ has an associated isotropy group, namely the set of elements of $G$ that fixes $x$,
\[
G_x=\{g\in G: g(x)=x\} \;.
\]
The conjugacy class of $G_x$ is called the orbit type and gives a stratification of $M$ into types. This induces a stratification of the orbit space $M/G$ with a smooth structure \cite{Pflaum01}. When the group action is free (or slightly more generally, if the isotropy group is the same for each point) the orbit space is a manifold, and $M$ has the structure of a principal bundle over the orbit space. This is the standard setting for symmetry reduction. Reconstruction, i.e., the study of the dynamics in the fiber, may still be challenging globally because the fiber bundle may not be trivial: it may not have a global section.
Many interesting group actions are not free. When the Lie group is compact, singular reduction is still possible, where ``singular'' reminds us that the reduced space in this case is not a manifold. The standard tool is the Hilbert basis of invariants of the group action, which together with all relations and inequalities give an accurate description of the singular reduced space, see e.g.~\cite{Chossat02,Pflaum01}. We will use this approach in passing in some of the examples in \Sec{Examples} and \Sec{Comparison}.
The principal orbit bundle (or principal stratum) is the stratum that is open and dense in $M$ \cite{Pflaum01,Chossat00,Golubitsky88}. Moreover, its quotient by $G$ is connected \cite[Thm 4.3.2]{Pflaum01}. In many cases, the isotropy group of the principal orbits is the trivial group. Within the principal stratum there is a subgroup of $G$ that acts freely, so we return to the simpler case of a free action, albeit on the smaller space obtained by removing certain closed subsets from the original $M$. The fact that the symmetry commutes with the map $f$ implies that every stratum of $M$ is a forward invariant set of $f$. When $f$ is invertible, this set is both forward and backward invariant, but for non-invertible maps there may be points outside
the invariant set that map into it.
Hence when $f$ is a diffeomorphism, the principal stratum is an open subset of full dimension that is invariant under $f$. In the examples to follow, the ``global Poincar\'e section" of the symmetry is often taken to be a section on this principal stratum.
For example, consider the circle action on ${\mathbb{ C}}^2$ given by
\beq{circleAction}
G = \{\psi_\tau( z_1, z_2) = ( e^{il \tau} z_1, e^{im \tau } z_2)|\; (z_1,z_2) \in {\mathbb{ C}}^2,\; \tau \in {\mathbb{ S}}^1\}
\end{equation}
for positive integers $l$ and $m$. Note that if $\psi_t$ is a symmetry of a map $f$, then
$f$ must have, e.g., the invariant $\Im( \bar z_1^m z_2^l)$, where $ \Im(z)$ represents the imaginary part of $z$.
When $l=m=1$ the only non-trivial isotropy group is found at the origin. The principal stratum hence is the open set ${\mathbb{ C}}^2 \setminus \{0\}$, on which the action of $\psi_\tau$ is free, so that it becomes a bundle with fiber ${\mathbb{ S}}^1$. Nevertheless, the principal stratum ${\mathbb{ C}}^2 \setminus \{ 0\}$ does not have a global Poincar\'{e} section. This would imply that it is a fiber bundle with base ${\mathbb{ S}}^1$, which is impossible since every loop in ${\mathbb{ C}}^2 \setminus \{ 0\}$ is contractible.
The action may be restricted to the invariant hyperboloids given by the level sets $\Im( z_1 \bar z_2) = L$. Whenever $L \not = 0$, the hyperboloid topologically is ${\mathbb{ R}}^2 \times {\mathbb{ S}}^1$ and there is a global section. Consequently upon removing the whole cone $\Im( z_1 \bar z_2) = 0$ from ${\mathbb{ C}}^2$, each connected component does have a global section. Alternatively one may remove one circle from each invariant ${\mathbb{ S}}^3$ defined by $|z_1|^2 + |z_2|^2 = const$, which turns ${\mathbb{ S}}^3$ into ${\mathbb{ R}}^2 \times{\mathbb{ S}}^1$. Which ``surgery'' to chose depends on properties of the map $f$.
When $l=1$ and $m>1$ there is a stratum in addition to the origin given by $(0, z_2)$ with discrete isotropy group $t = j / m$, $j = 1, \dots, m-1$. Every map with symmetry $\psi_\tau$ has this set as an invariant plane. Removing this plane gives the principal stratum $A \times {\mathbb{ C}}$, where $A = {\mathbb{ C}} \setminus \{ 0 \}$. The section $\Im(z_1) = 0$, $\Re(z_1) > 0$ has the set $z_1 = 0$ as an invariant boundary, and hence this section is a global Poincar\'{e} section for maps with symmetry $\psi$. Thus we find a trivial bundle $A \times {\mathbb{ C}}$ with fiber ${\mathbb{ S}}^1$ and base ${\mathbb{ R}}^+ \times {\mathbb{ C}}$.
When both $l$ and $m$ are bigger than 1 there are three strata: the origin $(0,0)$, the plane $(0, z_2)$ and the plan $(z_1, 0)$. Removing these from ${\mathbb{ C}}^2$ turns the manifold into $A^2$ where $A = {\mathbb{ C}} \setminus \{ 0 \}$. A section along the positive real axis in either punctured plane $A$ gives a global Poincar\'{e} section.
\section{Examples}\label{sec:Examples}
In this section we will give several examples to illustrate \Th{Reduction}.
We start with a classical two-dimensional example.
\bexam[(M\"obius Map)]
An elliptic M\"obius transformation is conjugate to the form
\[
f(z) = \frac{\phantom{+}az+b}{-bz+a} \;,
\]
for real $a,b$ with $ab \neq 0$.
This map has fixed points at $z = \pm i$ and is analytic on the upper half plane $M = \{z\in {\mathbb{ C}}: \Im(z) > 0\}$.
As was noted by \cite{Maeda87}, $f$ has the symmetry $Y = (z^2+1)\frac\partial{\partial z}$, with flow
\[
\psi_t(z) = \frac{\cos(t)z + \sin(t)}{\cos(t)-\sin(t)z} \;.
\]
On $M$ the orbits of $\psi$ are simply circles of radius $r$ and center $i\sqrt{r^2+1}$ for any $r \ge 0$. The action represented by the symmetry flow is not free, since $z=i$ is a fixed point of $\psi_t$. However on the principal stratum $M \setminus\{i\}$, the action is free, and has the global Poincar\'{e} section $\Sigma = \{i\sigma: 0<\sigma<1\}$, with return time $T = 2\pi$.
Notice that in order to achieve a global Poincar\'{e} section, the map the fixed point must be removed from the upper half plane.
To compute the reduced form \Eq{skewProduct}, we use the covering \Eq{covering}, $p(\sigma,\tau) = \psi_\tau(\sigma)$ and the notation \Eq{skewProduct}, $F(\sigma,0) = (k(\sigma),\rot(\sigma))$ on the section, to obtain
\begin{align*}
f \circ p(\sigma,0) &= f(\sigma)=
\frac{a\sigma +b}{a-b\sigma}\\
p \circ F(\sigma,0) &= \psi_{\rot(\sigma)}(k(\sigma)) =
\frac{\cos(\rot(\sigma))\,k(\sigma) + \sin(\rot(\sigma))}{\cos(\rot(\sigma))-\sin(\rot(\sigma))\,k(\sigma)} \;.
\end{align*}
Since, by \Eq{liftDef}, these must be equal we conclude that $F$ takes the form
\[
F(\sigma,\tau) = (\sigma, \tau + \rot(\sigma)) \;,
\]
where $\rot(\sigma)\equiv\rot$ is, in fact, any constant such that $\rot = \arg( a + i b)$.
The most natural choice may be the principal value $\operatorname{Arg}( a + i b) \in (-\pi, \pi]$, and each branch of $\arg$ gives a different lift.
Note that both $f$ and $Y$ have the invariant
\[
I(z) = \frac{1}{\Im(z)} (|z|^2+1) \;.
\]
According to \Lem{Invariant}, an invariant for $k$ is naturally $\iota^*I = \sigma + \frac{1}{\sigma}$, or trivially $\sigma$ itself.
\end{example}
\bexam[(Twisted Symmetries):]
On the manifold $M={\mathbb{ R}}^2\times ({\mathbb{ R}}/{\mathbb{ Z}})$, define
\beq{moebius}
f(\xi,z)= \left( R_{\beta z}\,h\left( R_{-\beta z} \xi \right),\, z
+\alpha(|\xi|) \right) \;,
\end{equation}
where $\alpha: {\mathbb{ R}}^+ \to {\mathbb{ R}}$, $R_\theta \in SO(2)$ is the rotation by angle $2\pi \theta$, and $\beta \in {\mathbb{ R}}$. The map $h: {\mathbb{ R}}^2 \to {\mathbb{ R}}^2$ is smooth and, so that $f$ be continuous on $M$, is assumed to have the symmetry
\beq{twistedMap}
h\circ R_\beta = R_\beta \circ h\;.
\end{equation}
Note that if $h$ were to commute with $R_\theta$ for all $\theta$, then $R_{\beta z} h(R_{-\beta z} \xi) = h(\xi)$, so \Eq{moebius} would already be written in the skew-product form. Instead we assume that $\beta$ is rational, so that the symmetry \Eq{twistedMap} is discrete.
The map \Eq{moebius} has the symmetry
\beq{rotationalFlow}
\psi_t(\xi,z)=(R_{\beta t}\,\xi ,z+t) \;.
\end{equation}
This flow has a global Poincar\'{e} section $\Sigma = \{(\sigma,0): \sigma \in {\mathbb{ R}}^2\}$, with the covering map $p(\sigma,\tau) = \psi_\tau(\sigma,0)$.
Since $\Sigma$ is simply connected, \Th{Reduction} implies there exits a lift $F$ which can be easily computed using \Eq{kDefine}:
\beq{twistedLift}
F(\sigma,\tau) = (R_{-\beta\alpha(|\sigma|) }\,h(\sigma), \tau + \alpha(|\sigma|)) \;.
\end{equation}
Since $\beta = \frac{p}{q}$ is rational, then by \Eq{twistedMap} $h$ must have the associated \emph{discrete} $(p,q)$ symmetry. In this case, the orbits of $\psi$ are $(p,q)$ torus knots on $M \setminus \{\xi = 0\}$. The action of the symmetry group is not free since the isotropy group of any point on the $z$-axis is different from those not on the axis. This example shows that even when the orbit space $M/G$ is not a manifold, the reduced map of \Th{Reduction} may still be as smooth as the original map. The point is that \Eq{twistedLift} has a residual discrete symmetry with the action $g(\xi,\tau) = (R_\beta\xi,\tau)$. Factoring out this discrete symmetry would indeed lead to a singular reduced space.
An example of an orbit of the reduced map in the section $\Sigma$ is shown in \Fig{sandDollar}, and many other examples of such maps can be constructed by replacing $h$ with any map with a discrete symmetry like those in \cite{Field09}.
\InsertFig{sandDollar}{Points on an attractor for the reduced map $k$ in \Eq{twistedLift} for $\beta = \frac{1}{3}$. Here, using the complex form $u = x+iy$, following \cite{Chossat88}, we took $h(u,\bar u) = (\lambda + u \bar u)u + \gamma \bar u ^2$.
The figure shows the case $\lambda = -2.43$, $\gamma = 0.1$ and $\alpha(|\sigma|) = - 2.67$.}{sandDollar}{3in}
\end{example}
\bexam[(Non-orientable manifold):]
For $(\xi, z) \in {\mathbb{ R}}^2 \times {\mathbb{ R}}$, let $G$ be the discrete group generated by $g(\xi,z)=(-\xi,z+1)$. Note that $G \simeq {\mathbb{ Z}}$, and the action $(g,(\xi,z))\mapsto g(\xi,z)$ on ${\mathbb{ R}}^3$ is free and proper. This implies that $M=\mathbb{R}^3/G$
is a manifold; in this case $M$ is non-orientable.
Let $f:\mathbb{R}^3\to\mathbb{R}^3$ be the map
\beq{moebius2}
f(\xi,z)= \left( R_{(z +\beta)/2}h\left( R_{-z/2}\xi \right),\, z +\beta \right).
\end{equation}
where (as in the previous example) $R_\theta$ is the rotation by angle $2 \pi \theta$, $h: {\mathbb{ R}}^2 \to {\mathbb{ R}}^2$, and $\beta\in\mathbb{R}$. Note that since $R_{(z+1)/2} = -R_{z/2}$, $f \circ g = g \circ f$, thus $f$ can be thought of as a map on the quotient $M$.
It is easy to see that the flow
\[
\psi_t(\xi,z)=(R_{t/2} \xi,z+t) \;.
\]
is a symmetry of \Eq{moebius2}. Note that $\psi_{t}\circ g=g\circ\psi_{t}$ so that $\psi_t$ defines a flow on $M$. Moreover, since $\psi_{t+1}=g\circ\psi_t$ the orbits of the symmetry $\psi$ are embedded circles in $M$, and $\psi$ has the global Poincar\'{e} section
\[
\Sigma=\{(\sigma,0): \sigma \in {\mathbb{ R}}^2\} \subset M\;.
\]
The return map for this case is simply $r_\psi(\sigma) = \psi_1(\sigma) = -\sigma$. Computation of the reduced map using \Eq{kDefine} gives $F(\sigma,\tau) = (h(\sigma), \tau + \beta)$.
Note that \Eq{moebius2} is volume preserving on $M$ whenever $h$ is area preserving; for example, the map
\[
h(x,y)= \left(y, \frac{2a y}{y^2+1}-x\right)
\]
is area preserving for any $a \in {\mathbb{ R}}$. This map also has the integral
\[
J(x,y)=y^2 x^2+x^2-2a y x+y^2 \;,
\]
i.e., one has $J\circ h= J$. In this case the function
\[
I(\xi,z)=J\left(R_{-z/2} \xi\right) \;.
\]
is an integral of $f$. Since it is also an integral of $\psi_t$, $I\circ \psi_t=I$, \Lem{Invariant} implies that $h$ has the reduced integral $\iota^* I = J$, as we already knew.
\end{example}
\bexam[(Nonhyperbolic Cat Maps):]
Suppose $f$ is a diffeomorphism of the $n$-torus $M={\mathbb{ R}}^n/{\mathbb{ Z}}^n$ that fixes the origin, $f(0) = 0$, and is homotopic to the map $b(\xi)=B\xi$, where $B\in SL(n,{\mathbb{ Z}})$ is an $n\times n$ unimodular matrix.
We will assume that $B$ has a simple eigenvalue $\lambda = 1$ with right eigenvector $v$
and left eigenvector $w$. The vector field $V = v \cdot \nabla$ generates the flow
\beq{CatFlow}
\psi_t(\xi)=\xi+t\,v \;,
\end{equation}
In order that $V$ be a symmetry of $f$, condition \Eq{MapSymmetry} requires that $Df(\xi)v=v$
for all $\xi\in M$. For example, the map
\beq{HomoCat}
f(\xi)=B\xi+\phi(\xi) v \;,
\end{equation}
with $\phi(0) = 0$, is of the assumed form. It has the symmetry $\psi_t$ if $\phi(\xi+t\,v)=\phi(\xi)$, for all $t$, so that $D\phi(\xi) v = 0$.
Since $w$ is a left eigenvector of $B$, the set $\Sigma = \{\xi \in M : w \cdot \xi = 0 \mod 1\}$ is a codimension-one subspace invariant under $b$ and transverse to $v$. Indeed since $B$ has integer coefficients, $\Sigma\hookrightarrow M$ is a codimension-one torus; an example is sketched in \Fig{3torus}. Moreover, it is clear that since $\Sigma$ is transverse to $v$, it is a Poincar\'e section for the flow \Eq{CatFlow}.
Since $f$ and $b$ are homotopic and both fix the origin, $f_*=b_*$, as maps on the fundamental group ${\pi_1}(M,0)$.
Moreover, $\Sigma$ is an invariant set for $b$, so the group $\iota_*{\pi_1}\left(\Sigma,0\right)$ is invariant under
$b_*$. Consequently,
\[
f_*\iota_*{\pi_1}\left(\Sigma,0\right)=b_*\iota_*{\pi_1}\left(\Sigma,0\right)
\subseteq \iota_*{\pi_1}\left(\Sigma,0\right) \;.
\]
Therefore, \Th{Reduction} and condition \Eq{newSubset} imply that there exists a lift $F$
of $f$ to the cover $\Sigma\times{\mathbb{ R}}$ of the form of \Eq{skewProduct} with a covering map $p:\Sigma\times{\mathbb{ R}}\to M$ of the form $p(\sigma,\tau) = \sigma + \tau v$. To simplify the computations we will use an equivalent cover,
$\widetilde{p}:{\mathbb{ T}}^{n-1}\times {\mathbb{ R}}\to M$ given by
\[
\widetilde{p}(\sigma,\tau) = S \sigma + \tau v\;,
\]
where $(S\,|\,v) \in SL(n,{\mathbb{ Z}})$ and the columns of the $n \times(n-1)$ integer matrix $S$ form a basis for $\Sigma$. Consequently,
\beq{Bhat}
B (S\,|\,v) = (S\,|\,v) \begin{pmatrix} \hat B & 0 \\ 0 & 1 \end{pmatrix}
\end{equation}
with $\hat B \in SL(n-1,{\mathbb{ Z}})$.
In this case, a lift of the form \Eq{skewProduct} exists and must satisfy \Eq{liftDef}. For example, a lift of \Eq{HomoCat} is
\[
F(\sigma,\tau) = (\hat B \sigma, \tau +\omega(\sigma) ) \;,
\]
where $\omega(\sigma)=\phi(S\sigma)$.
\InsertFig{3torus.pdf}{The global Poincar\'{e} section $\Sigma$ of the flow \Eq{CatFlow} for the matrix \Eq{CatMap} is an embedded submanifold of the three-torus $M$.}{3torus}{2.4in}
An explicit case of the form \Eq{HomoCat} for $n=3$ is
\beq{CatMap}
B=\begin{pmatrix} 0 & 1 & 0 \\
0 & 0 & 1 \\
1 & -4 & 4
\end{pmatrix} \;,
\quad \mbox{and } \phi(x,y,z) = g(x-y,y-z)
\end{equation}
for a function $g: {\mathbb{ T}}^2 \to {\mathbb{ R}}$ such that $g(0)=0$.
Note that $1$ is an eigenvalue of $B$ with right and left eigenvectors $v=(1,1,1)$ and $w = (-1,3,-1)$, respectively.
Thus the two-torus
\beq{catSection}
\Sigma=\{(x,y,z)\in M: x-3y+z=0 \mod 1\} \;,
\end{equation}
shown in \Fig{3torus}, is invariant under $b$ and is a global Poincar\'{e} section for the flow \Eq{CatFlow}. Moreover, since $D\phi v = 0$, the map $f$ has symmetry \Eq{CatFlow}.
The fundamental group of $\Sigma$ is generated by the loops $\eta_1(t)=(2t,t,t)$ and $\eta_2(t)=(t,t,2t)$. Setting $u_i=\iota_*[\eta_i]$ for $i=1,2$, it is easy to verify that
$f_*u_1=u_2$ and $f_*u_2= -u_1+ 3u_2$. Consequently, \Th{Reduction} implies there exists a lift of $f$ of the form \Eq{skewProduct}. Indeed, the covering map
\[
\widetilde{p}(\sigma,\tau) = \begin{pmatrix}
2 & -1 & 1 \\ 1 & -1 & 1 \\ 1 & -2 & 1\end{pmatrix}
\begin{pmatrix} \sigma_1 \\ \sigma_2 \\ \tau\end{pmatrix} \;,
\]
gives $(\sigma_1, \sigma_2, \tau) = (x-y,y-z,-x + 3y-z)$, so that $\tau = 0$ corresponds to $\iota(\Sigma)$. Upon computing $\hat B$ from \Eq{Bhat}, the lift takes the form \Eq{skewProduct}
with
\[
\begin{split}
k(\sigma) &= \begin{pmatrix} 0 & 1 \\ -1 & 3 \end{pmatrix} \sigma \;,\\
\rot(\sigma) &= g(\sigma_1,\sigma_2) \;.
\end{split}
\]
Since the eigenvalues of $k$ are $\gamma^{\pm 2}$ where $\gamma$ is the golden mean, the reduced map is an Anosov diffeomorphism.
Note that the one cannot freely replace \Eq{catSection} with any global section of \Eq{CatFlow} and still satisfy the topological requirement \Eq{lift}. For example, the torus $\tilde \Sigma = \{(0,y,z) \in M\}$ is also a global section for \Eq{CatFlow}. However, the map $b(\xi) = B\xi$ takes the generators $[(0,t,0)]$ and $[(0,0,t)]$ of the fundamental group of $\tilde \Sigma$ into $[(t,0,-4t)]$ and $[(0,t,4t)]$, violating \Eq{lift}.
\end{example}
\bexam[(Resonant Circle Action):]
The flow $\psi_\tau(z_1,z_2) = (e^{il \tau} z_1, e^{i m \tau} z_2)$ of \Eq{circleAction}
corresponds to a circle action on $M = {\mathbb{ C}}^2$ and is familiar from the study of resonant, coupled oscillators.
As already noted in \Sec{OrbitSpaces} the action \Eq{circleAction} is not free: it has a fixed point at $(0,0)$ and points with nontrivial, discrete isotropy for each point $(z_1, 0)$ whenever $l>1$ and for each point $(0, z_2)$ whenever $m > 1$.
The general map on ${\mathbb{ C}}^2$ with the symmetry \Eq{circleAction} can be written
(see, e.g., \cite{Golubitsky88})
\[ \begin{split}
z'_1 = f_1(\rho) z_1 + f_2(\rho)\bar z_1^{m-1} z_2^l \;, \\
z'_2 = f_3(\rho) z_2 + f_4(\rho) z_1^m \bar z_2^{l-1} \;.
\end{split}\]
where the $f_j(\rho)$ are complex-valued functions of the real invariants of \Eq{circleAction}, namely
\beq{HilbertBasis}
\begin{array}{ll}
\rho_1 = z_1 \bar z_1 \;, & \rho_2 = z_2 \bar z_2 \;, \\
\rho_3 = \Re( z_1^m \bar z_2^l) \;, & \rho_4 = \Im( z_1^m \bar z_2^l) \;. \\
\end{array}
\end{equation}
These four invariants form a \emph{Hilbert basis}: every function invariant under the action \Eq{circleAction} is a function of $\{\rho_1,\rho_2,\rho_3,\rho_4\}$ (see also \Sec{Comparison}). For example, $\det{Df}$ is invariant under the symmetry $\psi_\tau$, and hence a function of the invariants.
As a particular example consider
\[
(z'_1, z'_2) = ( z_1 - i \varepsilon m \bar z_1^{m-1} z_2^l,
z_2 - i \varepsilon l z_1^m \bar z_2^{l-1}) \;.
\]
This map is, to lowest order in $\varepsilon$, the time-$\varepsilon$ flow of the Hamiltonian $H =\rho_3$, and so is approximately a four-dimensional symplectic map in $(x_1,x_2,y_1,y_2)$, up to terms of order $\varepsilon^2$. The flow of this Hamiltonian has two invariants, namely $H$ itself, and the Hamiltonian that generates the symmetry, $l \rho_1 + m \rho_2$. When $\varepsilon \ll 1$ these functions will be approximate invariants of the map as well.
Assuming, for concreteness, that $m > l \ge 1$, then since $\psi_t$ does not act freely on $M={\mathbb{ C}}^2$, there are points with nontrivial isotropy, in particular the set $\{(0, z_2): z_2 \in {\mathbb{ C}}\}$. This set is forward invariant under both $\psi_t$ and $f$, and
removing it from $M$ gives the manifold $\tilde M = A \times {\mathbb{ C}} \simeq {\mathbb{ R}}^+ \times{\mathbb{ S}}^1 \times {\mathbb{ C}}$. On this manifold $\psi_t$ has the global Poincar\'e section
\[
\Sigma = \{(x_1,z_2), x_1> 0, z_2 \in {\mathbb{ C}}\} \subset \tilde M \;.
\]
Thus to get a global section, it is not necessary to restrict to the principal stratum, which would entail, when $l>1$, the removal of the points $(z_1,0)$ as well. On the section $\Sigma$ these additional points are fixed under $f$.
The restriction of $f$ to $\tilde M$ has a lift on $\Sigma \times {\mathbb{ R}}$ determined by \Eq{kDefine}, i.e., by the requirement
\[
e^{-i\rot} (x_1 - i \varepsilon m x_1^{m-1} z_2^l,
z_2 - i \varepsilon l x_1^m z_2^{l-1}) \in \Sigma \;.
\]
Thus $\rot$ is determined by requiring the first component to be real and positive, and $F$ is given by
\[\begin{split}
k(x_1, z_2) &= ( x_1 |1 - i \varepsilon m x_1^{m-2} z_2^l |,
e^{-i \rot} (z_2 - i \varepsilon l x_1^m z_2^{l-1} ) ) \;,\\
\rot(x_1,z_2) &= \arg( 1- i \varepsilon m x_1^{m-2} z_2^l) \;.
\end{split}\]
Note that even though the set $z_1 = 0$ is forward invariant, since $f$ is not invertible
there are points that map into the invariant set.
Similarly, the reduced map $k$ has points that map into the boundary $x_1 = 0$
given by the solutions of $z_2^l = ( i \varepsilon m x_1^{m-2})^{-1} $. The reduced map $k$ can be extended to the excluded set $x_1 =0$ by continuity, so that the pre-images
of $x_1 = 0$ have well defined orbits. Even so, the fibre map $\rot$ is undefined for points that map into the set $x_1 = 0$ so, strictly speaking, reduction by lifting fails in this case. However, even though the fibre map is undefined for certain points, the reduced map $k$ is well behaved.
\end{example}
\section{Volume-Preserving Symmetry Reduction} \label{sec:VP}
We will now show that, when the map $f$ and its symmetry are both volume preserving, the reduced map $k$ of \Eq{skewProduct} is also volume preserving on $\Sigma$, with respect to an appropriate volume form.
This specializes \Th{Reduction} to the volume-preserving setting.
We denote the volume form by $\Omega \in {\Lambda}^n(M)$; by assumption, both $f$ and $\psi_t$ preserve this form, $f^*\Omega = \Omega$ and $\psi_t^* \Omega = \Omega$, respectively. Equivalently, the symmetry vector field is incompressible: $L_Y \Omega = (\nabla \cdot Y) \Omega = 0$, where $\nabla \cdot Y$ is the divergence of $Y$. When $\nabla \cdot Y=0$,
we will say that $Y$ is incompressible.
\begin{thm}\label{thm:ReducedVolumeForm} In addition to the hypotheses of \Th{Reduction}, assume that $f$ is volume preserving and its symmetry $Y$ is incompressible. Then the reduced map $k$ of \Eq{skewProduct} preserves a volume form $\nu$ on $\Sigma$ defined by
\beq{nuDefn}
\nu = \iota^*i_Y \Omega \;.
\end{equation}
\end{thm}
\begin{proof}
First we note that $\mu := i_Y \Omega$ is an $(n-1)$-form on $M$. Moreover, since $\Sigma$ is a section, $Y \not\in T\Sigma$, and so $\nu = \iota^*\mu$ is non-degenerate on $\Sigma$. In addition,
\(
d\nu = \iota^* di_Y\Omega = \iota^*(L_Y\Omega-i_Yd\Omega) = 0 \;;
\)
thus $\nu$ is a volume form. Finally, from \Eq{MapSymmetry} we have
\(
f^*\mu = i_{f^*Y} f^*\Omega = i_Y \Omega = \mu \;,
\)
so $f$ preserves $\mu$.
We now assert that, in fact,
\beq{form-aux}
\Pi^*\nu = p^*\mu \;,
\end{equation}
where $\Pi:\Sigma\times {\mathbb{ R}}\to \Sigma$ is the canonical projection.
To see this recall that since $p$ is a local diffeomorphism and $Y\neq0$ everywhere,
there exists a flow box \cite{Abraham78} in $M$ near each point of $\Sigma$.
Therefore, we can reduce the proof to the case in which $\Sigma$ is an open set in
${\mathbb{ R}}^{n-1}$, $M$ is of the form $M=\Sigma\times{\mathbb{ R}}$
and $p$ is the identity. If we let $(\sigma,\tau)$ be the coordinates of $M$
then, by construction, we have that $Y=\partial/\partial\tau$ and
the flow on $M$ is $\psi_t(\sigma,\tau)=(\sigma,\tau+t)$. In these coordinates,
we have that $\iota\circ\Pi(\sigma,\tau)=(\sigma,0)$, and the volume form
can be written as
\[
\Omega=\kappa(\sigma,\tau)d\sigma_1\wedge\cdots\wedge d\sigma_{n-1}\wedge d\tau \;.
\]
Since $\psi_t$ is volume preserving, we conclude that
$\kappa(\sigma,\tau)=\kappa(\sigma,0)$, for all $(\sigma,\tau)\in M$.
This implies that $\mu=i_Y \Omega $ does not depend on $\tau$ and
therefore
\(
p^*\mu =(\iota\circ\Pi)^*\mu=\Pi^*\nu \;,
\)
which is equality \Eq{form-aux}.
Finally from the definition \Eq{reducedMap} of $k$, and the fact that $\Pi\circ j=id_\Sigma$,
we get that
\[
k^*\nu=j^* F^*\Pi^*\nu=j^* F^*p^*\mu
=j^*p^*f^*\mu=j^*p^*\mu=j^*\Pi^*\nu=\nu \;.
\]
In this way, we conclude that $k$ preserves $\nu$.
\end{proof}
Theorem \ref{thm:ReducedVolumeForm} can be combined with \Cor{FlowReduction} to show that $\nu$ is a reduced volume form for an incompressible vector field with an incompressible symmetry.
\begin{cor}\label{cor:ReducedVolumeForm}
Suppose that $X,Y \in \mathcal{V}(M)$ are incompressible, commuting vector fields and that $Y$ has a flow $\psi$ with a global Poincar\'{e} section $\Sigma$ that is an immersed manifold $\iota:\Sigma\hookrightarrow M$. Then there exists a vector field
$K$ on $\Sigma$ and a function $\zeta: \Sigma \to {\mathbb{ R}}$, such that
\beq{decomposition}
\left.X\right|_\Sigma=K+ \zeta \left.Y\right|_\Sigma \;,
\end{equation}
where $K$ is incompressible with respect to the volume form $\nu=\iota^*i_Y\Omega$ on $\Sigma$.
\end{cor}
\begin{proof}
By \Th{Reduction} the flow $\varphi$ of $X$ has a lift $\Phi$ to the cover $\Sigma\times {\mathbb{ R}}$ of the form \Eq{FlowReduction}. Let $K \in \mathcal{V}(\Sigma)$ denote the vector field generated by the reduced flow $k$ of $\Phi$. By
\Th{ReducedVolumeForm}, each $k_t$ preserves the volume-form $\nu$ so that
$K$ is incompressible on $\Sigma$ with respect to the form in \Eq{nuDefn}.
Finally, since $K$ is tangent to---and $Y$ is transverse to---$\Sigma$, the
vector field $X|_\Sigma$ can be written as a linear combination of $K$ and $Y$.
Indeed \Eq{FlowReduction} gives
\[
\left.\frac{\partial}{\partial t}\right|_{t=0}\Phi_t(\sigma,\tau)=
K + \left.\frac{\partial}{\partial t}\right|_{t=0} \omega(\sigma,t)\frac{\partial}{\partial \tau} \;.
\]
Since $Y=\frac{\partial}{\partial \tau}$ in the coordinates $(\sigma,\tau)$, this implies \Eq{decomposition} with $\zeta(\sigma)= \dot\omega(\sigma,0)$.
\end{proof}
A special case of this corollary is included in a theorem of Haller and Mezic \cite{Haller98}, who treated the case $M = {\mathbb{ R}}^3$ and noted that the flow on the two-dimensional group orbit space is Hamiltonian.
\bexam[(Volume-Preserving normal form):]
Consider the family of volume-preserving diffeomorphisms, $f: {\mathbb{ T}}^3 \to {\mathbb{ T}}^3$, defined by
\beq{VPNormal}
f(x,y,z) = (x+a(y), y+z+b(y),z+b(y)) \;.
\end{equation}
Similar maps on ${\mathbb{ R}}^3$,
arise as the normal form for a volume-preserving map near saddle-node bifurcation with a triple-one multiplier \cite{Dullin08a}. The map \Eq{VPNormal} has symmetry $Y = \frac\partial{\partial x}$ which generates the flow $\psi_t(x,y,z) = (x+t,y,z)$.
The natural section for $\psi$ is $\Sigma = \{(0,v,w): v,w\in {\mathbb{ T}}^2\}$, and the corresponding covering map $p: {\mathbb{ T}}^2 \times {\mathbb{ R}} \to {\mathbb{ T}}^3$ defined by $p(v,w,\tau) = \psi_\tau(0,v,w)$. The group $\iota_*{\pi_1}(\Sigma,0)$ is generated by the equivalence classes $\gamma_1 = [(0,t,0)]$ and $\gamma_2 = [(0,0,t)]$.
Since $a$ and $b$ are a periodic functions,
\[
f_*(\gamma_1) = \gamma_1 \;, \quad f_*(\gamma_2) = \gamma_1 + \gamma_2 \;,
\]
If we assume that $a(0)=b(0)=0$, $f$ fixes the origin, and then the requirement \Eq{newSubset} holds.
Of course \Eq{VPNormal} is already in skew-product form, with $\tau = x$, and $\rot(v,w) = a(v)$, and the reduced map becomes $k(v,w) = (v+w+q(v), w+q(v))$. The latter is a generalized Chirikov standard map and is typically chaotic.
Notice that if $a(y) = y$ the topological condition would fail for any choice of section. One way to see this is to note that the linear part of the map would then be a Jordan block that does not have any $2$-dimensional invariant subspaces. Another is to note that lifting with the natural section would give $\rot(u,v) = u$, which is not a continuous function from ${\mathbb{ T}}^2 \to {\mathbb{ R}}$. Finally, one might think that this problem could be repaired by first lifting $f$ to the universal cover ${\mathbb{ R}}^3$, eliminating the topological requirement. However, in this case almost all orbits of the linear map are unbounded--- there is no return map to any surface.
\end{example}
\bexam[(4D Volume-preserving map):]
Consider a map $f = T_2 \circ T_1$ of ${\mathbb{ R}}^4 \simeq {\mathbb{ C}}^2$ with coordinates $(z_1, z_2, \bar z_1, \bar z_2)$
given by the composition of the two ``shears''
\begin{align*}
T_1( z_1, z_2) = \left( z_1( a + i \sqrt{ c_1 - a^2 +
h_1(z_2 \bar z_2)/(z_1 \bar z_1)}), z_2 \right) \\
T_2( z_1, z_2) = \left( z_1, z_2( b + i \sqrt{ c_2 - b^2 +
h_2(z_1 \bar z_1)/(z_2 \bar z_2)}) \right),
\end{align*}
where $a,b,c_1,c_2$ are real constants, and $h_1$ and $h_2$ are non-negative real functions.
Each of the shears has constant Jacobian $c_i$, so $\det Df = c_1 c_2$. Thus if
$c_1c_2 = 1$, $f$ is volume preserving.
If $c_1 = c_2 =1$ the map $f$ is the composition of two symplectic maps.
Each shear commutes with the two symmetries $\psi_t(z_i) = e^{it} z_i$. Taken together this implies that $f$ has a pair of commuting symmetries, i.e., is equivariant under the ${\mathbb{ T}}^2$ action $\psi_{(t_1, t_2)}( z_1, z_2) = ( e^{it_1} z_1 , e^{i t_2} z_2)$.
The reduction of \Th{ReducedVolumeForm} can be carried out recursively, thus reducing $f$ to an area-preserving map.
To construct a global section, we must remove the sets of nontrivial isotropy from $M$: removing the fixed sets $z_1 = 0$ and $z_2 = 0$ leaves $A^2 = ({\mathbb{ C}} \setminus \{0\})^2$. On this set there is a global Poincar\'e section for the two-parameter action, $\Sigma = \{(x_1,x_2): x_i > 0\} \subset A^2$.
The result of the reduction is to simply introduce polar coordinates in both complex planes, and give a fiber map with $-\rot = ( \arg( z_1') , \arg( z_2') )$.
The reduced map $k$ would be written in terms of $(x_1, x_2) \in \Sigma$.
Instead we use the two invariant radii $\rho_1 = z_1 \bar z_1$ and $\rho_2 = z_2 \bar z_2$
as coordinates we find
\[
T_1( \rho_1, \rho_2) =( c_1 \rho_1 + h_1(\rho_2)), \rho_2) = (\rho_1', \rho_2)
\]
and similarly for $T_2$, so that the reduced area-preserving map (assuming $c_1 c_2 = 1$) is
\[
T_2(T_1(\rho_1,\rho_2)) = (c_1\rho_1+ h_1(\rho_2), c_2\rho_2 + h_2(\rho_1'))\;,
\]
which is smooth even for $\rho_j = 0$.
\end{example}
\section{Noether symmetries}\label{sec:Invariant}
For Hamiltonian flows and symplectic maps the close relationship between the existence of invariants and symmetries is well-known. Noether's theorem, for example, implies that a Lagrangian system with a symmetry acting in configuration space has an invariant. This relationship is exploited in the Liouville-Arnold construction of action-angle variables for an integrable $n$-degree-of-freedom Hamiltonian where each invariant generates a symmetry vector field. In this case, the involutive property of the invariants implies that the corresponding symmetry vector fields commute.
As we remarked in the introduction, analogues of Noether's theorem do not exist for maps in general nor for volume-preserving maps in particular. For example the skew product
\[
f(x,y,\tau) = (k(x,y),\tau + \rot(x,y))
\]
on $\Sigma \times {\mathbb{ S}}^1$ has the obvious symmetry $Y =\partial/\partial\tau$ and preserves the volume form $\Omega = dx \wedge dy \wedge d\tau$ whenever $k$ is area preserving. If $k$ has no invariant, then neither does $f$. This would occur, e.g., when $\Sigma = {\mathbb{ T}}^2$, and $k(x,y) = (2x+y,x+y)$, the cat map.
In addition, a three-dimensional map could have an invariant, restricting its orbits to two-dimensional level sets, and yet have no symmetry: the map can be chaotic on the level sets. Examples include the trace maps \cite{Roberts94}, such as the Fibonacci map
\[
f(x,y,z) = (y,z,-x+2yz) \;.
\]
Trace maps are volume preserving; many similar examples have also been constructed \cite{Gomez02}.
Consequently, symmetries do not necessarily give rise to invariants without some additional structure.
One sufficient additional structure corresponds to symplectic maps with Hamiltonian symmetries.
A symmetry $Y$ that is a Hamiltonian vector field is known as a \emph{Noether symmetry}.
Here we give a slight generalization of a theorem of \cite[App. 1]{Bazzani88},
which required the symplectic map to be a twist map. We show that one can drop this requirement if the map has a mild recurrence property.
\begin{thm}[Symplectic Noether]\label{thm:SymplecticNoether}
If $Y$ is a Noether symmetry of a symplectic map $f$ and $f$ has any recurrent orbits, then $f$ has an invariant. Conversely if $I$ is an invariant of a symplectic map $f$, then its Hamiltonian vector field is a symmetry of $f$.
\end{thm}
\begin{proof}
For this proof let $\mu$ denote the symplectic form (e.g., $\mu = dq \wedge dp$), and let $I$ be the Hamiltonian of the vector field $Y$: $i_Y\mu = dI$. Consequently
\(
f^* i_{Y} \mu = d ( f^*I)
\),
but by the symmetry and symplectic conditions
\(
f^* i_{Y} \mu = i_{f^*Y} f^*\mu = i_{Y} \mu = dI
\).
Consequently $d(f^*I-I) = 0$, so that
$
I(f(x)) = I(x) +c
$
for some constant $c$. Suppose $x^*$ is a recurrent point, then there exists a subsequence $t_i \to \infty$ such that $f^{t_i}(x^*) \to x^*$. However, since
\(
I(f^t(x)) = I(f^{t-1}(x))+c = \ldots = I(x) + tc
\),
this implies that
\[
\lim_{i\to \infty} [I(f^{t_i}(x^*) - I(x^*)] = \lim_{i \to \infty} t_i c = 0 \;,
\]
so that $c = 0$. Thus $I$ is an invariant for $f$.
Conversely, if the symplectic map $f$ has an invariant $I$, then it generates a Hamiltonian vector field $Y$, and
\[
i_{f^*Y} \mu = f^*(i_Y \mu) = f^*(dI) = d(f^*I) = dI = i_Y \mu \;.
\]
Thus $i_{f^*Y -Y} \mu =0$, but since $\mu$ is non-degenerate, this can only occur when $f^*Y = Y$.
\end{proof}
The condition that $f$ has a recurrent orbit is necessary. The map
\[
f(x,y) = (x+c, y + g(x)) \;,
\]
on ${\mathbb{ R}}^2$ is symplectic with the canonical form $dx \wedge dy$ and has the translation symmetry $Y =\partial/\partial y$. This symmetry is generated by the Hamiltonian $I(x,y) = x$, which, however, is not an invariant for $f$ for any $c \neq 0$. Note that $f$ has no recurrent orbits.
By contrast, if we take the same $f$ but suppose that it is a map on ${\mathbb{ S}}^1 \times {\mathbb{ R}}$, then $f$ may have recurrent orbits, but the symmetry vector field $\partial/\partial y$ is locally, but not globally Hamiltonian, and so does not generate an invariant.
\bexam[(Lyness Map):]
For any $a > 0$, the Lyness map \cite{Lyness45},
\beq{Lyness}
f(x,y) = \left(y, \frac{a+y}{x}\right)
\end{equation}
is a diffeomorphism on the positive quadrant $M = {\mathbb{ R}}^2_+$ that
preserves the symplectic form $\mu = \frac{1}{xy} dx \wedge dy$.
It has the symmetry \cite{Beukers98}
\[
Y = {xy} ( -\partial_y I, \partial_x I)
=-xy\frac{\partial I}{\partial y}\frac{\partial }{\partial x}
+xy\frac{\partial I}{\partial x}\frac{\partial }{\partial y} \;,
\]
where
\[
I = \frac{(1+x)(1+y)(a+x+y)}{xy} \;.
\]
The function $I$ is an invariant for $f$.
Moreover, $Y$ is the Hamiltonian vector field of $I$ with
respect to $\mu$ and so $I$ is an invariant for $Y$. This is
in agreement with \Th{SymplecticNoether}.
The level sets of the invariant $I$ for $I > I_{min} = I(x^*,x^*)$ are topologically circles that intersect the line
$\Sigma = \{(\sigma,\sigma): \sigma > x^*\}$ with $x^* = \frac12(1+\sqrt{1+4a})$ exactly once. The symmetry is not free since the fixed point $(x^*,x^*)$ has nontrivial isotropy. However on the principal stratum $\tilde M = M \setminus\{(x^*,x^*)\}$, $\Sigma$ is a global Poincar\'e section for the flow of $Y$.
Given that $\Sigma$ is simply connected, $f$ can be reduced with a lift.
The lifted map on $\Sigma \times {\mathbb{ R}}$
is $F(\sigma,\tau) = (\sigma, \tau + \rot(\sigma))$.
An integral expression for $\rot(\sigma)$ can be obtained from the results in \cite{Beukers98}.
\end{example}
In the next section, we will apply these results to a four-dimensional symplectic map.
\section{Comparison of reduction methods} \label{sec:Comparison}
In this section, we will illustrate three different approaches to reduction, namely
\begin{itemize}\setlength{\itemsep}{0mm}
\item invariants,
\item polar coordinates, and
\item reduction by lifting.
\end{itemize}
We will use, as an example, a four-dimensional symplectic map with a rotational symmetry.
A symplectic map on $M={\mathbb{ R}}^4 \simeq {\mathbb{ C}}^2$ of standard type with a symmetry can be constructed using the Lagrangian generating function
$S(z,z') = \frac12 |z - z'|^2 - W(z \bar z)$ with $z = x + i y \in {\mathbb{ C}}$ and $W:{\mathbb{ R}}\to{\mathbb{ R}}$ a $C^2-$function.
This generating function is invariant under the rotation $( e^{i \tau} z, e^{i \tau} z')$.
Introducing the conjugate momenta $w = p_x + i p_y$, then the corresponding 4D map, generated
by $w = -\partial_{\bar z} S$ and $w' = \partial_{\bar z'} S$, is
\beq{4DSymplectic}
f(z,w) = ( z' , w') = \left( z + w', w - z V( z \bar z) \right)
\end{equation}
where $V(x)=\frac{d}{dx}W(x)$.
The map \Eq{4DSymplectic} has the symmetry $\psi_t(z, w) = ( e^{it} z, e^{it} w)$, e.g., the flow \Eq{circleAction} with $k=l=1$. Note that $\psi_t$ is symplectic and is the flow of the ``angular momentum" Hamiltonian, $L = \Im( \bar z w)$. The discrete Noether theorem, \Th{SymplecticNoether}, implies that $L$ is an invariant of $f$ as well.
\paragraph*{Reduction using invariants:}
Since the symmetry $\psi_t$ is a special case of \Eq{circleAction}, the basis \Eq{HilbertBasis} with $(z_1,z_2) \to (z,w)$ and $k=l=1$ generates its real polynomial invariants.
Note that these $\rho_i$ are not independent: they obey the relation
\beq{relation}
\rho_3^2 + \rho_4^2 = \rho_1 \rho_2 \;,
\end{equation}
and the inequalities
\[
\rho_1 \ge 0 \;,\quad \rho_2 \ge 0 \;.
\]
The map $(z, w) \to (\rho_1,\rho_2,\rho_3,\rho_4)$ is called the \emph{Hilbert map}.
Rewriting \Eq{4DSymplectic} in terms of the invariants gives
\bsplit{Hilbert}
\rho_1' &= \rho_1( 1 - V)^2 + \rho_2 + 2 \rho_3(1 - V) \;, \\
\rho_2' &= \rho_2 - 2 \rho_3 V + \rho_1 V^2 \;,\\
\rho_3' &= \rho_3 - \rho_1 V + \rho_2' \;,
\end{split}\end{equation}
where $V= V(\rho_1)$
and $\rho_4 = -L$, the negative of the constant angular momentum. This is volume preserving with respect to the form $\Omega = d\rho_1 \wedge d\rho_2 \wedge d\rho_3$ on ${\mathbb{ R}}^3 [ \rho_1, \rho_2, \rho_3]$ and has an invariant $\rho_4^2 = \rho_1\rho_2-\rho_3^2$ inherited from the relation \Eq{relation}.
The Hilbert coordinates satisfy the Poisson structure given by
$\{\rho_1, \rho_2 \} =4 \rho_3$, $\{ \rho_1, \rho_3 \} = 2\rho_1$, $\{ \rho_2, \rho_3\} = - 2\rho_2$ with Casimir $\rho_1 \rho_2 - \rho_3^2$.
Moreover, the reduced mapping \Eq{Hilbert} is a Poisson map: it preserves this reduced Poisson structure.
The level set of the Casimir is a smooth submanifold provided $\rho_4 \not = 0$. On each such a level set (a smooth symplectic leaf) the map is area preserving. The reduction by the symmetry and by its generating invariant reduces the dimension of the map by two.
However, the invariant set $\rho_4 = 0$ is a half-cone. This case is a singular reduction, because the action $\psi_t$ is not free, since the origin $(z, w) = (0,0)$ is a fixed point for the whole group.
From this construction, the map in the symmetry direction, in
the form $\tau' = \tau + \rot(\rho_i)$, can be obtained as follows. The Hilbert map from full phase space $T^*{\mathbb{ R}}^2$ to the invariants is not invertible. It has the symmetry orbits as fibers. As pre-image in the fiber we can choose
\[
(z, w) =h(\rho) =: \left(\sqrt{\rho_1}, \sqrt{\rho_2}
e^{i \arg(\rho_3 - i\rho_4)}\right) \;.
\]
The angle increment $\rot$ is now defined by $f(h(\rho)) = \psi_{\rot} h( \rho')$.
The explicit form of the fiber map therefore is found from $z' = e^{i \rot} \rho_1'$
so that $\rot = \arg( z')$. The map $h$ is not defined when $\rho_3 = \rho_4 = 0$, i.e.\ at the origin of the $z$-plane and at the origin of the $w$-plane.
\paragraph*{Reduction using polar coordinates:}
Polar coordinates appear whenever the symmetry angle $\theta$ becomes a coordinate. There is a unique extension of this transformation of coordinates to a symplectic transformation of all of phase space that is linear in the momenta, the so-called cotangent lift, see, e.g.~\cite{Abraham78}. Denoting the new variables by $(r, \theta, p_r, p_\theta)$, the momentum $p_\theta$, conjugate to $\theta$, becomes the conserved quantity.
Specifically set $z = r e^{i\theta}$ with cotangent lift $w = e^{i\theta} ( p_r - i p_\theta/r)$ and rewrite the map in the new coordinates $(r, \theta, p_r) = ( \sqrt{ z \bar z}, \arg z, \Re( z \bar w )/r)$.
Having done most of the work in for the Hilbert basis already, we proceed as follows. In terms of the invariants, the polar coordinates are $(r, p_r, p_\theta) = ( \sqrt{\rho_1}, \rho_3/\sqrt{\rho_1}, \rho_4 )$ and the map is obtained from that of the invariants by eliminating $\rho_2$ using $\rho_2 = (\rho_3^2 + \rho_4^2)/\rho_1 = p_r^2 + p_\theta^2/r^2$. This gives an area-preserving map that depends on the parameter $p_\theta$
\bsplit{PolarCoord}
r' &= \left( p_\theta^2 / r^2 + (r + p_r - r V)^2 \right)^{1/2} \;, \\
p_r' &= \left( r'^2 - r( r + p_r - r V)\right) / r' \;,
\end{split}\end{equation}
where $V = V(r^2)$. The fiber map is obtained from rewriting $z' = z + w' = z + w - z V$ in polar coordinates,
hence $r' e^{i\theta'} = e^{i\theta}( r + p_r - i p_\theta/r - r V)$ and therefore
\[
\theta' = \theta + \arg( r + p_r - r V - i p_\theta/r) \;.
\]
This map works well unless $r' = 0$, which makes $p'_r$ and $\theta'$ undefined.
Points with $r'=0$ can only be reached when $p_\theta = 0$, but in that case it happens
as soon as $r + p_r = r V$. Excluding $p_\theta = 0$ gives a well-defined, smooth map.
\paragraph*{Reduction by lifting:}
For reduction by lifting we need to find a global Poincar\'{e} section of the symmetry
flow $\psi_t$. As a first attempt try the surface defined by $\Im(z) = 0$ for $x =\Re(z) > 0$. The flow is tangent to the section when $\Re(z) = 0$, hence all of the plane $\{z = 0\}$ is tangent to the section. Unfortunately, this plane of tangency is not
invariant under the map. In order to avoid this difficulty, we can choose a larger
invariant set that contains the tangency set, and restrict the map to the complement of
this set.
The angular momentum $L(z,w)=-\Im( z \bar w)=\Im( \bar z w)$ is invariant both under the flow and the symplectic map. Notice that the plane $\{z = 0\}$
is contained in the level set $\{L(z,w)=0\}$, that is invariant.
The manifold $M \setminus \{ L(z,w) = 0\}$ has two connected components,
\[
\tilde M=\{(z,w)\in{\mathbb{ C}}\times{\mathbb{ C}}: L(z,w)>0 \} \;,
\]
and the corresponding negative angular momentum set. As the analysis for either component is the same, we will now restrict the map $f$ and the flow $\psi_t$ to $\tilde M$.
The symmetry flow as a global Poincar\'{e} section on $\tilde M$:
\[
\Sigma=\{(x,w)\in \tilde M: x>0\} \;.
\]
This section is relatively closed in $\tilde M$ and if $(x,w)\in\Sigma$ then $\Im(w)>0$.
We know that the lift of $f$ is of the form
$F(\sigma,\tau)=(k(\sigma),\tau+\rot(\sigma))$. Taking a point $\sigma=(x, w) \in \Sigma$
we can compute its image under \Eq{4DSymplectic} to obtain
\beq{LiftReduced}
f(x,w)=(x + w - x V(x^2), w - x V(x^2)) \;.
\end{equation}
By \Eq{kDefine}, this equals $\psi_{\rot(\sigma)}(k(\sigma))$, so that $e^{-i \rot} ( x + w - x V)$ must be real and positive (where $x$ is assumed positive). Noticing that $\Im( x + w - x V)>0$, this immediately gives $\rot(\sigma) = \arg( x + w - x V)$, and we can choose the branch so that $0<\rot(\sigma)<\pi$. The reduced map thus becomes
\[
k(x,w) = ( |x + w - x V| , e^{-i \rot(x,w)} (w - x V)) \;.
\]
By \Eq{nuDefn}, $k$ preserves the volume form $\nu = p_y dx \wedge dp_x \wedge dp_y$.
Because we started with a symplectic map the reduced map has an invariant, $\iota^* \Im(\bar z w) = x \Im( w) = x p_y = p_\theta$. If we were to eliminate $p_y$ by fixing the invariant, we would again obtain the polar coordinate map \Eq{PolarCoord}.
\paragraph*{Comparison:}
The Hilbert map is similar to what we call the projection $\Pi$ to the section $\Sigma$,
while its ``inverse'' $h$ is like the inclusion map $\iota$. In differential geometry, a fiber bundle is said to have a ``global section'' if $h$ is continuous.
Notice, however, that the Hilbert map has fiber ${\mathbb{ S}}^1$, while when we construct the
global Poincar\'{e} section we have a bundle with base ${\mathbb{ S}}^1$ and fiber $\Sigma$.
In fact, as pointed out in \Sec{CircleAction}, some finite covering of the bundle is a direct product with ${\mathbb{ S}}^1$. In the present example the manifold $\tilde M$ is
topologically a direct product ${\mathbb{ S}}^1 \times {\mathbb{ R}} \times {\mathbb{ R}}^+$.
The Hilbert map may be applied to points with isotropy where projection to $\Sigma$ is not defined. The reduced map \Eq{Hilbert} is well-defined, even when the ${\mathbb{ S}}^1$-bundle does not possess a global section or when there is no global Poincar\'{e} section.
Reduction by lifting is more efficient in that one does not need to introduce extra coordinates; moreover, one does not need to know the ``good'' coordinate system to do the reduction.
Instead the choice of global Poincar\'{e} section defines such a coordinate system.
\section{Conclusions}\label{sec:Conclusions}
We have studied maps with continuous symmetries, specializing to the case that the symmetries have a global Poincar\'e section. We showed in \Th{Reduction} that if a necessary topological condition is satisfied, then the map has a lift such that in certain coordinates the lift has the skew-product form \Eq{skewProduct}. We called this ``reduction by lifting" because the map on the section, the reduced map $k$, describes the dynamics modulo the symmetry. The fiber map, which corresponds to a translation, is obtained naturally from the symmetry flow.
The topological conditions for the existence of a lift require that the homotopy homomorphism induced by the map leaves the fundamental group of the Poincar\'e section invariant. This requirement is trivial if the section is simply connected. In principle, the restriction could be avoided if one first lifts the dynamics to the universal cover; however---as we saw in the examples---almost all orbits may become unbounded. In many cases, a minimal lift using only the symmetry flow seems more useful.
If the map and its symmetry are volume-preserving, then the reduced map is also volume-preserving as shown in \Th{ReducedVolumeForm}. If both the map and its symmetry preserve an invariant, then the reduced map also has an invariant. Finally, if the map and its symmetry are symplectic, then, as in Noether's theorem, the existence of a symmetry implies an invariant, providing the map has recurrent orbits, recall \Th{SymplecticNoether}.
A number of examples were given to compare and contrast reduction by lifting to standard reduction techniques. Reduction by lifting is more parsimonious than the Hilbert mapping technique, which can result in a high-dimensional map that must satisfy a number of constraints. It is more explicit than the standard reduction by group orbit technique, and moreover applies when the action of the symmetry is neither proper nor free. When the action is not free, reduction by lifting need not lead to a map on a singular space.
However, the existence of a global section for the symmetry flow is a strong restriction. This can be overcome in some cases, as we showed, by restriction of the dynamics to invariant strata of the symmetry flow.
|
\section{Introduction}
Recently, it has been recognized that entanglement does not depict
all possible quantum correlations contained in a bipartite state.
Based on the measurement on the subsystem in the bipartite system, quantum
discord was proposed as a measure for the quantum correlation beyond
entanglement~\cite{Zurek00,Ollivier02,Vedral02,Celeri11,Modi2011}. Quantum
discord was viewed as a figure of merit for characterizing the nonclassical
resources in the deterministic quantum computation with one-qubit~\cite{Knill98,Datta08}.
It was also discussed that zero-discord of the initial system-environment
states is a necessary and sufficient condition for completely positivity
of reduced dynamical maps~\cite{Rosario08,Shabani09}. At the same
time, a necessary and sufficient condition for nonzero-discord was
also given for any dimensional bipartite states~\cite{Vedral11}.
Most recently, operational interpretations of quantum discord were
proposed in~\cite{Datta11,Winter11}, where quantum discord was shown
to be a quantitative measure about the performance in the quantum
state merging~\cite{Horodecki05}. Over the past decade, quantum
discord has received a lot of attentions in Refs.~\cite{Cen11,Modi10,Lu11,Luo10,Adesso11,Guzi11,Fei11,Chen11,
RioNature,Ferraro10,Ali10,Streltsov11,Piani11,Cornelio11,Fanchini11a,Mazzola10,Lanyon08},
and also see the reviews \cite{Modi2011,Celeri11}.
In general, quantum discord is upper bounded by the entropy of the
measured subsystem~\cite{Datta1003,Xi11,LuoPRAa,LuoPRAb,Terno11},
but it has remained an open question when this bound is saturated
for general mixed states. For this question, some sufficient conditions
were given in~\cite{Xi11,LuoPRAa,LuoPRAb,Zhang11}. This motivates
us to systematically investigate the upper bound of quantum discord
and give a necessary and sufficient condition for saturating the upper
bound of quantum discord. On the other hand, as an achievable upper
bound of the quantum discord, the von Neumann entropy of a system
can be considered as the quantum correlative capacity which tells
us how strongly can this system be quantum correlated with others~\cite{LuoPRAa}.
So we want to ask another question: what can the closeness between the quantum
discord and the upper bound tell us?
It is interesting that there are monogamic relations between different
measures on correlations~\cite{Koashi04}. In this paper, by purifying
the bipartite systems and using the Koashi-Winter relation~\cite{Koashi04},
we obtain a new monogamic relation: a system being quantum correlated
with another one limits its possible classical correlation with a
third system. For the bipartite systems, the total amount of quantum
discord between two subsystems and the classical correlation between
the measured subsystem and the environment cannot exceed the von Neumann
entropy of the measured subsystem. We further prove that the necessary and sufficient condition for saturating the upper bound of quantum
discord is equivalent to the equality condition for the Araki-Lieb
inequality~\cite{AL70,Nielsen,LR73a,LR73b}. And we give the explicit
characterization of the quantum states saturating the upper bound
of quantum discord, through which we demonstrate that saturating the
upper bound of quantum discord means that the measured subsystem cannot
be further correlated with the environment. For a two-qubit system,
quantum discord is strictly less than the von Neumann entropy
of the measured qubit of two-qubit states other than when the two-qubit
system in pure states.
This paper is organized as follows. In Sec. \ref{sec:review QD},
we give a brief review on quantum discord. In Sec. \ref{sec:UP_QD},
we show that there is a trade-off between the quantum discord and
the classical correlation, and then we prove a necessary and sufficient condition for the saturating of the upper bound of quantum
discord. Section \ref{sec:conclusion} is the conclusion.
\section{Review of quantum discord}
\label{sec:review QD}
First, we recall the concepts and properties of the quantum discord.
Let Hilbert space $\mathcal{H}=\mathcal{H}_{A}\otimes\mathcal{H}_{B}$
be a bipartite quantum system. Let $\mathcal{D}(\mathcal{H})$ be
a set of bounded, positive-semidefinite operators with unit trace
on $\mathcal{H}$. Given a bipartite quantum state $\rho^{AB}\in\mathcal{D}(\mathcal{H})$,
the von Neumann mutual information between the two subsystems $A$
and $B$ is defined as~\cite{Nielsen}
\begin{equation}
I(A:B):=S(A)+S(B)-S(AB),\label{eq:TC}
\end{equation}
where $S(X):=-\mathrm{Tr}\rho^{X}\log_{2}\rho^{X}$ is
the von Neumann entropy, $\rho^{X}$ is a quantum state of system $X$~\cite{Nielsen}. The mutual information $I(A:B)$
is a measure of the total amount of correlations in the bipartite
quantum state. Generally speaking, it was divided into quantum correlation
and classical correlation~\cite{Vedral02,Vedral03,Zurek00,Ollivier02}.
The classical correlation was seen as the amount of information about
the subsystem $A$ that can be obtained via performing a measurement
on the other subsystem $B$. Then, the measure of the classical correlation~\cite{Vedral02}
is defined by
\begin{equation}
J(A|B):=\max_{\{E_{i}^{B}\}}\Big(S(A)-\sum_{i}p_{i}S(A|E_{i}^{B})\Big),\label{eq:CC}
\end{equation}
where $\{E_{i}^{B}\}$ is the positive operator valued measure (POVM)
on $B$, $S(A|E_{i}^{B})$ is the von Neumann entropy of the post-measurement
states $\rho_{i}^{A}=$Tr$_{B}(E_{i}^{B}\rho^{AB})$ corresponds to
outcome $i$ with the probability $p_{i}=$Tr$(E_{i}^{B}\rho^{AB})$.
Therefore, quantum discord~\cite{Zurek00,Ollivier02} is defined by the
difference of the quantum mutual information and the classical correlation
\begin{equation}
D(A|B):=I(A:B)-J(A|B).\label{eq:QDA_B}
\end{equation}
Quantum discord is asymmetric with respect to $A$ and $B$,
in general, $D(A|B)\neq D(B|A)$, and it is always nonnegative~\cite{Zurek00,Ollivier02}. Quantum discord
vanishes if and only if there exists an optimal choice of measurement
$\{E_{i}^{B}\}$ on $B$ leaves the state $\rho^{AB}$ unperturbed~\cite{Terno11}.
The condition for zero-discord can been reduced to the equality condition
using relative entropy~\cite{Hayden04,Datta1003}, and was also discussed
in Refs~\cite{Ollivier02,Vedral11,Datta1003}. For any bipartite
pure state $|\psi\rangle^{AB}$, one checks that
\begin{equation}
D(B|A)=S(A)=S(B)=D(A|B).
\end{equation}
\section{The upper bound of quantum discord}
\label{sec:UP_QD}
\subsection{A trade-off between quantum discord and classical correlation}
For any general bipartite mixed state $\rho^{AB}$, we can always
find a tripartite pure state $\rho^{ABE}=|\Psi\rangle^{ABE}\langle\Psi|$
such that $\rho^{AB}=\mathrm{Tr}_{E}(\rho^{ABE})$, where $E$ represents
the environment. Hereafter, we will consider this purification about
general bipartite mixed state $\rho^{AB}$. The monogamic relation
between the entanglement of formation and the classical correlation
between the two subsystems is given by~\cite{Koashi04} \begin{subequations}
\begin{align}
E_{F}(A:E)+J(A|B)=S(A),\label{eq:KW2}
\end{align}
where $E_{F}(A:E)$ is entanglement of formation (EoF), defined as
$E_{F}(A:E)=\min_{\{p_{i},|\psi_{i}\rangle\}}\sum_{i}p_{i}S(\mathrm{Tr}_{E}(|\psi_{i}\rangle\langle\psi_{i}|))$,
where the minimum is taken over all pure ensembles $\{p_{i},|\psi_{i}\rangle\}$
satisfying $\rho^{AE}=\sum_{i}p_{i}|\psi_{i}\rangle\langle\psi_{i}|$~\cite{Bennett96,Wootters98,Wootters01}.
This relation is universal for any tripartite pure states. Thus, we
can give an other reorder version of Koashi-Winter relation,
\begin{align}
E_{F}(E:A)+J(E|B)=S(E).\label{eq:KW6}
\end{align}
\end{subequations} Due to the symmetric property of the EoF, i.e.,
$E_{F}(A:E)=E_{F}(E:A)$, we can eliminate the EoF by combining Eqs.~(\ref{eq:KW2})
and (\ref{eq:KW6}), then we obtain
\begin{align}
J(A|B)-J(E|B)=S(A)-S(E).
\end{align}
To be clearer, we substitute $D(A|B)=I(A:B)-J(A|B)$ into the above
equation, and obtain a trade-off between the quantum discord and the classical correlation
as follows
\begin{equation}
D(A|B)+J(E|B)=S(B).\label{eq:KWX_1}
\end{equation}
This new monogamic relation tell us that the amount of quantum correlation
between $A$ and $B$, plus the amount of classical correlation between
$B$ and the complementary part $E$, must be equal to the entropy
of the measured subsystem $B$. More importantly, based on this new
monogamic equation, we can introduce
\begin{equation}
\tilde{J}(B/A):=S(B)-D(A|B),
\end{equation}
which quantifies classically correlative capacity of $B$ with other
systems except $A$. In other words, for general tripartite mixed
states $\rho^{ABC}$, the classical correlation between $B$ and $C$
cannot be greater than $\tilde{J}(B/A)$. To be convinced, let us
purify $\rho^{ABC}$ as $\rho^{ABC}=\mathrm{Tr}_{E}|\Psi\rangle^{ABCE}\langle\Psi|$,
then we have $J(CE|B)=\tilde{J}(B/A)$ due to the monogamic relation
(\ref{eq:KWX_1}). Because the classical correlation is non-increasing
under the local quantum operation~\cite{Vedral02}, then we have
\begin{equation}
J(C|B)\leq J(CE|B)=\tilde{J}(B/A),
\end{equation}
which is equivalent to
\begin{equation}
D(A|B)+J(C|B)\le S(B).\label{eq:tradeoff_mixed_states}
\end{equation}
We emphasize that Eq.~(\ref{eq:tradeoff_mixed_states}) is applicable
for any tripartite state $\rho^{ABC}$ and the equality holds if $\rho^{ABC}$
is pure. We can see that the total amount of the quantum discord between
two subsystems and the classical correlation between the measured
subsystem and the environment cannot exceed the von Neumann entropy
of the measured subsystem.
\subsection{The upper bound of quantum discord}
The monogamic relation (\ref{eq:KWX_1}) directly supplies a general
upper bound for the quantum discord, which was proved in Refs.~\cite{Datta1003,Xi11}.
In the following, we are going to determine which states saturate
this bound. Combining the monogamic relation (\ref{eq:KWX_1}) with
the equality condition for Araki-Lieb inequality, we have the following
result.
\begin{theorem}
\label{theorem:1} For the bipartite state $\rho^{AB}$, we have
\begin{equation}
D(A|B)\leq S(B)\label{eq:qd_leq_SB}
\end{equation}
with equality if and only if there exist a decomposition of $\mathcal{H}^{A}$
as $\mathcal{H}^{A^{L}}\otimes\mathcal{H}^{A^{R}}$ such that
\begin{equation}
\rho^{AB}=\rho^{A^{L}}\otimes|\psi\rangle^{A^{R}B}\langle\psi|.
\end{equation}
\end{theorem}
To prove this theorem, we first introduce a lemma as follows.
\begin{lemma} \label{lemma:1}Considering different correlations
and entropy in the bipartite states, the following conditions are
equivalent
\begin{enumerate}
\item $D(A|B)=S(B)$;\label{enu:equivalence1}
\item $S(A)-S(B)=S(AB)$;\label{enu:equivalence2}
\item $E_{F}(A:B)=S(B)$.\label{enu:equivalence3}
\end{enumerate}
If one of them is satisfied, then the others are satisfied.
\end{lemma}
\begin{proof}
The equivalence of these conditions can be proved by the trade-off
relation (\ref{eq:KWX_1}) for the purification $|\psi^{ABE}\rangle$.
From Eq. (\ref{eq:KWX_1}), we can see that $D(A|B)=S(B)$ is equivalent
to $J(E|B)=0$. It is known that the classical correlation vanishes
if and only is the states are product states, see Refs.~\cite{Vedral02,Vedral03}.
So we have
\begin{eqnarray}
J(E|B)=0 & \Leftrightarrow & \rho^{EB}=\rho^{E}\otimes\rho^{B}\label{eq:equavalence_EB_product}\\
& \Leftrightarrow & S(EB)=S(B)+S(E).\label{eq:equivalence2_0}
\end{eqnarray}
For the tripartite pure states, we have $S(A)=S(EB)$ and $S(E)=S(AB)$.
Thus Eq.(\ref{eq:equivalence2_0}) is equivalent to $S(A)-S(B)=S(AB)$
and we obtain the equivalence between the first and second conditions.
Meantime, the third condition is equivalent to $J(B|E)=0$ due to
the Koashi-Winter relation $E_{F}(A:B)=S(B)-J(B|E)$. Again using
the property of the classical correlation, we can see the third condition
is equivalent to $\rho^{EB}=\rho^{E}\otimes\rho^{B}$, and in turn
equivalent to the first condition.
\end{proof}
Form the above lemma, we will complete the proof of Theorem~\ref{theorem:1}.
\begin{proof}[Proof of Theorem~\ref{theorem:1}]
The inequality was recently given in~\cite{Xi11,Datta1003}, and
it is also a direct consequence of Eq.(\ref{eq:KWX_1}).
Therefore, we obtain that quantum discord $D(A|B)$ saturates the upper bound $S(B)$
if and only if $S(A)-S(B)=S(AB)$, which is the equality condition
of the Araki-Lieb inequality \cite{AL70}
\begin{equation}
\left|S(A)-S(B)\right|\leq S(AB)
\end{equation}
when $S(A)\geq S(B)$.
Recently, the quantum states saturated the
Araki-Lieb inequality was explicitly given by Zhang and Wu \cite{Zhang11},
by using the explicit characterization of quantum states that saturates
the strong subadditivity inequality \cite{Hayden04} and the relation
between the Araki-Lieb inequality and the strong subadditivity inequality
\cite{Nielsen}. From the result in Ref~\cite{Zhang11}, we know
that $S(A)-S(B)=S(AB)$ if and only if there exist a decomposition
$\mathcal{H}^{A}=\mathcal{H}^{A^{L}}\otimes\mathcal{H}^{A^{R}}$ such
that
\begin{equation}
\rho^{AB}=\rho^{A^{L}}\otimes|\varphi\rangle^{A^{R}B}\langle\varphi|.
\end{equation}
This completes this proof of Theorem~\ref{theorem:1}.
\end{proof}
\begin{figure}
\centering{}\includegraphics[scale=0.8]{fig1} \caption{Schematics diagram for the Theorem~\ref{theorem:1} and~\ref{theorem:2}.
In both subfigures, the right side of the solid line are the measured
subsystems B. In (a), the quantum discord saturates the general upper
bound given by the von Neumann entropy of the measured subsystem $B$
if and only if there exists a decomposition of $\mathcal{H}^{A}$
as $\mathcal{H}^{A^{L}}\otimes\mathcal{H}^{A^{R}}$ such that $\rho^{AB}=\rho^{A^{L}}\otimes|\varphi\rangle^{A^{R}B}\langle\varphi|$;
In (b), the quantum discord equals the von Neumann entropy of the
unmeasured subsystem $A$ if there exist a decomposition of $\mathcal{H}^{B}$
as $\mathcal{H}^{B^{L}}\otimes\mathcal{H}^{B^{R}}$ such that $\rho^{AB}=|\psi\rangle^{AB^{L}}\langle\psi|\otimes\rho^{B^{R}}$.
\label{fig:trap}}
\end{figure}
This result shows that quantum discord is equal to the measured system
entropy if and only if the equality in the Araki-Lieb inequality holds.
It is an interesting thing that the measured subsystem $B$ cannot
correlate with the environment $E$ if the quantum discord between
$A$ and $B$ is equal to the entropy of the measured subsystem. In
other words, there must exists an isolated pure subsystem enclosing
the measured subsystem in the Hilbert space $\mathcal{H}_{A}\otimes\mathcal{H}_{B}$
when $D(A|B)=S(B)$. One can also see that given the entropy of the
measured subsystem, the maximal quantum discord between $A$ and $B$
will forbid system $B$ from being correlated to other systems outside
this composite system.
Due to the asymmetric property of quantum discord, one may ask whether
or not Theorem~\ref{theorem:1} still holds if we consider
$S(A)$ instead of $S(B)$. In fact, a conjecture about the von Neumann
entropy and quantum discord was presented by Luo, Fu and Li in Ref.~\cite{Luo10},
namely,
\begin{equation}
D(A|B)\leq\min\{S(A),S(B)\}.
\end{equation}
Later, Li and Luo showed that the part $D(A|B)\leq S(A)$ fails in
general, but might be true for low-dimension systems~\cite{LuoPRAa}.
We only have the following sufficient condition for the situation of $D(A|B)=S(A)$.
\begin{theorem}\label{theorem:2} For the bipartite state $\rho^{AB}$,
we have
\begin{equation}
D(A|B)=S(A),
\end{equation}
if the equality $S(B)-S(A)=S(AB)$ is satisfied. \end{theorem}
\begin{proof} According to the equality condition for the Araki-Lieb
inequality, $S(B)-S(A)=S(AB)$ is equivalent to that there exists a decomposition
$\mathcal{H}^{B}$ as $\mathcal{H}^{B^{L}}\otimes\mathcal{H}^{B^{R}}$
such that
\begin{equation}
\rho^{AB}=|\psi\rangle^{AB^{L}}\langle\psi|\otimes\rho^{B^{R}}.\label{eq:decomposition_B}
\end{equation}
For this density matrix, we can get
\begin{equation}
D(A|B)=D(A|B^{L})=S(A),
\end{equation}
where $D(A|B^{L})$ is the quantum discord for the pure state $|\psi\rangle^{AB^{L}}$.
\end{proof}
Besides, applying the Theorem \ref{theorem:1} and Lemma \ref{lemma:1},
we can see that all the following quantities are equal:
\begin{align}
D(A|B)=D(B|A)=E_{F}(A:B)=S(A)=S(B^{L})\label{eq:all_equals}
\end{align}
for the quantum state (\ref{eq:decomposition_B}).
As an illustration, let us consider the following example
\begin{equation}
\rho^{AB}=\frac{1}{4}\left(\begin{array}{cccccccc}
1 & 0 & 0 & 0 & 0 & 0 & 1 & 0\\
0 & 1 & 0 & 0 & 0 & 0 & 0 & 1\\
0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\
0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\
0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\
0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\
1 & 0 & 0 & 0 & 0 & 0 & 1 & 0\\
0 & 1 & 0 & 0 & 0 & 0 & 0 & 1
\end{array}\right),
\end{equation}
where $\mathcal{H}_{A}=\mathbb{C}^{2}$ and $\mathcal{H}_{B}=\mathbb{C}^{4}$.
The reduced states can be obtained $\rho^{A}=\frac{I^{A}}{2}$ and
$\rho^{B}=\frac{I^{B}}{4}$, where $I^{A}$ and $I^{B}$ are identity
operators on $\mathcal{H}_{A}$ and $\mathcal{H}_{B}$, respectively.
After some calculations one obtains
\begin{equation}
S(AB)=1,\: S(A)=1,\: S(B)=2.
\end{equation}
Therefore, this state satisfies $S(B)-S(A)=S(AB)$. On the other hand,
one checks that $\mathcal{H}_{B}=\mathcal{H}_{B^{L}}\otimes\mathcal{H}_{B^{R}}$
with $\mathcal{H}_{B^{L}}=\mathcal{H}_{B^{R}}=\mathbb{C}^{2}$. Then,
$\rho^{AB}$ is in the form of Eq.~(\ref{eq:decomposition_B}) with
$|\psi\rangle^{AB^{L}}=\frac{1}{\sqrt{2}}\big(|00\rangle+|11\rangle\big)$
and $\rho^{B^{R}}=\frac{I^{B^{R}}}{2}$, where $I^{B^{R}}$ is identity
operator on $\mathcal{H}_{B^{R}}$. With the result of Eq.~(\ref{eq:all_equals}),
we have
\begin{equation}
D(A|B)=D(B|A)=E_{F}(A:B)=1
\end{equation}
for this quantum state.
Through Theorem~\ref{theorem:1} and Theorem~\ref{theorem:2}, we identify
two conditions to witness the trapping of correlation within a pure-state
subsystem, see Fig. \ref{fig:trap}. This opens a new way to investigate
whether the correlation between A and B is essentially the correlation
between smaller parts of them.
\subsection{Arbitrary two-qubit states}
Now, we will discuss the application of our results in two-qubit system.
A two-dimension Hilbert space cannot be decomposed any more. The only
possibility of $|S(A)-S(B)|=S(AB)$ for two-qubit states is that $\rho^{AB}$
is a pure state. Hence, its upper bound is not reachable except two-qubit
pure state~\cite{Xi11}, namely, we have
\begin{equation}
D(B|A)<S(A),D(A|B)<S(B),
\end{equation}
for any mixed two-qubit state.
\section{conclusion}
\label{sec:conclusion}
In this work, we have given a new monogamic relation between the quantum discord and the classical
correlation in terms of the Koashi-Winter relations.
Based on the equality conditions for the Araki-Lieb inequality, we have
given a necessary and sufficient condition for the saturating of the
upper bound of quantum discord. We have shown that the subsystem of
the bipartite system can not correlated with the other system if the
quantum discord of the bipartite system is equal to the entropy of
the measured subsystem. We showed that there are some mixed states
where quantum discord is equal to entanglement of formation. In particular,
for two-qubit state, its upper bound is not reachable except pure
state.
\section{acknowledgments}
The authors are very grateful to the referees for helpful comments and criticisms. We thank A. Datta for interesting discussions. Z. J. Xi is supported
by the Superior Dissertation Foundation of Shaanxi Normal University
(S2009YB03). X.-M. Lu is supported by National Research Foundation and
Ministry of Education, Singapore (Grant No. WBS: R-710-000-008-271). X. Wang is supported by NSFC with
Grants No.11025527, No.10874151, and No.10935010.
Y. M. Li is supported by NSFC with Grant No.60873119, and the Higher
School Doctoral Subject Foundation of Ministry of Education of China
with Grant No.200807180005.
|
\section{Introduction}
Recently, surface states in topological insulators have attracted a
lot of attention in the condensed-matter
community~\cite{Hasan2010}. Both in two-dimensional (e.g., HgTe) and
in three-dimensional (e.g., Bi$_2$Se$_3$) compounds with strong
spin-orbit interaction the topological phase has been demonstrated
experimentally~\cite{Konig2007,Hsieh2008,Xia2009a,Xia2009b}. Although
these compounds are insulating in the bulk (since they have an energy
gap between the conduction band and the valance band), their surface
states support topological gapless excitations. In the simplest case
these low-energy excitations of a strong three-dimensional topological
insulator can be described by a single Dirac cone at the center of the
two-dimensional Brillouin zone ($\Gamma$
point)~\cite{Konig2007,Xia2009a,Hsieh2009a,Roushan2009,Hsieh2009b,*HZhang2009}.
The corresponding Hamiltonian is given by~\cite{Fu2008}
\begin{equation}
\mathcal{H}_0=
%
%
\hbar v_{F} \vec{\sigma} \cdot \vec{k} - \mu I.
%
\label{eq:hamiltonian0}
\end{equation}
Here $\vec{\sigma}$ represents a vector whose three components are the
three Pauli spin matrices, $I$ represents a $2\times 2$ identity
matrix in spin space, $v_{F}$ is the Fermi velocity, and $\mu$ is the
chemical potential. The low-energy excitations of $\mathcal{H}_0$
[Eq.~(\ref{eq:hamiltonian0})] are topologically protected against
perturbations~\cite{TZhang2009}. This has prompted recent research on
the transport properties of surface Dirac fermions. For example, the
conductance and magnetotransport of Dirac fermions have been studied
in normal metal -- ferromagnet (NF), normal metal -- ferromagnetic --
normal metal (NFN) and arrays of NF junctions on the surface of a
topological insulator~\cite{Mondal2010a,Mondal2010b,Zhang2010},
suggesting the possibility of an engineered magnetic switch. An
anomalous magnetoresistance effect has been predicted in
ferromagnetic-ferromagnetic junctions~\cite{Yokoyama2010}. Also,
electron tunneling and magnetoresistance have been studied in
ferromagnetic -- normal metal -- ferromagnetic (FNF)
junctions~\cite{Wu2010,Salehi2011}, for which it has been predicted
that the conductance can be larger in the anti-parallel configuration
of the magnetizations of the two ferromagnetic regions than in the
parallel configuration. In addition, a large research effort has been
devoted to studying models which predict the existence of Majorana
fermion edge states at the interface between superconductors and
ferromagnets deposited on a topological
insulator~\cite{Fu2008,Akhmerov2009,Tanaka2009}.
In this article we investigate adiabatic quantum pumping of Dirac
fermions through edge states on the surface of a strong
three-dimensional topological insulator. Quantum pumping refers to a
transport mechanism in meso- and nanoscale devices by which a finite
dc current is generated in the absence of an applied bias by periodic
modulations of at least two system parameters (typically gate voltages
or magnetic fields)~\cite{Buttiker1994,Brouwer1998,Spivak1995}. In
order for electrical transport to be adiabatic, the period of the
oscillatory driving signals has to be much longer than the dwell time
$\tau_{\rm dwell}$ of the electrons in the system, $ T = 2 \pi
\omega^{-1} \gg \tau_{\rm dwell} $. In the last decade, many different
aspects of quantum pumping have been theoretically investigated in a
diverse range of nanodevices, for example charge and spin pumping in
quantum dots~\cite{Switkes1999,Mucciolo2002,Sharma2003,Watson2003},
the role of electron-electron
interactions~\cite{Splettstoesser2005,Sela2006,Reckermann2010},
quantum pumping in graphene mono- and bilayers~\cite{Prada2009,
Zhu2009, Prada2010, Wakker2010, Tiwari2010, AlosPalop2011,
Kundu2011} as well as charge and spin pumping through edge states in
quantum Hall systems~\cite{miriam2003} and recently a two-dimensional
topological insulator~\cite{Citro2011}. On the experimental side,
Giazotto \textit{et al.}~\cite{Giazotto2011} have recently reported an
experimental demonstration of charge pumping in an InAs nanowire
embedded in a superconducting quantum interference device (SQUID).
Our main focus is to study quantum pumping induced by periodic
modulations of gate voltages or exchange fields, which are induced by
a ferromagnetic strip in two topological insulator devices: a NFN and
a FNF junction, see Figs.~\ref{fig:NFNjunction}
and~\ref{fig:FNFjunction}. Using a scattering matrix approach, we
obtain analytical expressions for the angle-dependent pumped current
in both types of junctions. We find that the adiabatically pumped
current in a NFN topological insulator junction induced by periodic
modulations of gate voltages reaches maximum values at specific energy
values. In order to explain the position of these values, we study in
detail the conductance of the junctions. In particular, we provide an
explanation for resonances in the conductance that were predicted but
not explained in detail in the previous
works~\cite{Mondal2010a,Mondal2010b}. We show that each time a new
resonant mode appears in the junction the conductance increases and
the pumped current reaches a maximum value. For the FNF pump we
predict a non-zero current by periodic modulation of the exchange
magnetic coupling in the absence of external voltages. We observe and
analyze basic similarities and differences between the two pumps
studied in this paper and highlight an experimentally distinguishable
feature between the pumped current and the conductance.
The remainder of the paper is organized as follows. In
Sec.~\ref{sec:NFN&FNF}, we describe the NFN and FNF junctions and use
a scattering matrix model to calculate the reflection and transmission
coefficients of both junctions. In Sec.~\ref{sec:conductance}, we
review the conductance of the NFN junction and present a detailed
analysis of the plateau-like steps that appear in the conductance. We
also analyze and compare the conductance of the FNF junction with
parallel and anti-parallel configuration of the magnetization. In
Sec.~\ref{sec:pumpedcurrent}, we calculate the adiabatically pumped
current for the two different pumps and derive analytical expressions
as a function of the angle of incidence of the carriers. We also
investigate the dependence of the pumped current and the conductance
on the width $d$ of the middle region. Finally, in
Sec.~\ref{sec:summary} we summarize our main results and propose
possibilities for experimental observation of our predictions.
\section{NFN and FNF junctions}
\label{sec:NFN&FNF}
\begin{figure}
\includegraphics[width=.8\columnwidth]{Plots/NFN.eps}
\caption{(Color online) Sketch of the N$_l$FN$_r$ junction on the
surface of a topological insulator. Pumping is induced by applying
gate voltages (not shown) to the normal leads. In the middle region
a thin ferromagnetic film induces ferromagnetism on the surface of
the topological insulator by means of the exchange
coupling~\cite{Yokoyama2010}. The arrow in the middle region
indicates the direction of the magnetization $M$ in this region.}
\label{fig:NFNjunction}
\end{figure}
We first describe the NFN junction, see
Fig.~\ref{fig:NFNjunction}. The junction is divided into three
regions: region $N_l$ (for $x < 0$), region $N_r$ (for $x > d$) and
the ferromagnetic region $F$ in the middle. The left and right-hand
side of the junction represent the bare topological insulator. The
charge carriers (surface Dirac fermions) in these regions are
described by the Hamiltonian $\mathcal{H}_0$
[Eq.~(\ref{eq:hamiltonian0})] whose eigenstates are given by
\begin{equation}
\psi_N^{\pm} = \frac{1}{\sqrt{2}}\, \left(
\begin{array}{c}
1 \\ \pm e^{\pm i \alpha}\\
\end{array} \right) e^{\pm ik_n x} e^{i q y} ,
\label{eq:eigenstates}
\end{equation}
where $+(-)$ labels the wavefunctions traveling from the left (right)
to the right (left) of the junction. The angle of incidence $\alpha$
and the momentum $k_n$ in the $x$-direction are given by:
\begin{equation}
\sin(\alpha) = \frac{\hbar v_F q}{|\epsilon + \mu|},
\label{eq:alpha}
\end{equation}
\begin{equation}
k_n = \sqrt{\left(\frac{\epsilon + \mu}{\hbar v_F}\right)^2 - q^2}.
\end{equation}
Here $\epsilon$ represents the energy measured from the Fermi energy
$\epsilon_F$ and $q$ denotes the momentum in the $y$-direction. In the
normal regions $N_l$ and $N_r$ a dc electrical voltage can be applied
via metallic top gates to tune the chemical potential $\mu$ and
thereby control the number of charge carriers incident on the
junction. We assume gate voltages to be small compared to the bandgap
for bulk states ($eV_i \ll E_g \sim 1$ eV, $i=l,r$), so that transport
is well described by surface Dirac states~\cite{HZhang2009}. In this
case, the eigenstates are given by
\begin{equation}
\psi_{N_l}^{\pm} = \frac{1}{\sqrt{2}}\, \left(
\begin{array}{c}
1 \\
\pm e^{\pm i \alpha_l}\\
\end{array} \right)
e^{\pm i k_{n_l} x }e^{i q y},
\label{eq:psiNl}
\end{equation}
\begin{equation}
\psi_{N_r}^{\pm} = \frac{1}{\sqrt{2}}\, \left(
\begin{array}{c}
1 \\
\pm e^{\pm i \alpha_r}\\
\end{array} \right)
e^{\pm i k_{n_r} (x-d) }e^{i q y},
\label{eq:psiNr}
\end{equation}
\begin{equation}
\sin(\alpha_i) = \frac{\hbar v_F q}{|\epsilon + \mu- eV_i|},
\label{eq:alphawithV}
\end{equation}
\begin{equation}
k_{n_i} = \sqrt{\left(\frac{\epsilon + \mu - eV_i}
{\hbar v_F}\right)^2 - q^2},
\end{equation}
where the index $i=l,r$ labels the normal sides of the junction.
In the middle region M of the junction ($0 < x < d$), the presence of
the ferromagnetic strip modifies the Hamiltonian by providing an
exchange field. The Hamiltonian that describes the surface states is now
$\mathcal{H} = \mathcal{H}_0 + \mathcal{H}_{\textrm{induced}}$, where
the induced exchange Hamiltonian is given by~\cite{Yokoyama2010,Mondal2010b}
\begin{equation}
\mathcal{H}_{\textrm{induced}} = \hbar v_F M \sigma_y,
\end{equation}
with the magnetization $\vec{M} = M\hat{y}$. The magnitude
$M$ depends on the strength of the exchange coupling of the
ferromagnetic film and can be tuned for soft ferromagnetic films by
applying an external magnetic field~\cite{Yokoyama2010}. The
eigenstates of the full Hamiltonian $\mathcal{H}$ are then given by:
\begin{equation}
\psi_F^{\pm} = \frac{1}{\sqrt{2}}\, \left(
\begin{array}{c}
1 \\
\pm e^{\pm i \alpha_m}\\
\end{array} \right)
e^{\pm i k_m x}e^{i q y} ,
\label{eq:psiF}
\end{equation}
with
\begin{equation}
\sin(\alpha_m) = \frac{\hbar v_F (q+M)}{|\epsilon + \mu|},
\label{eq:beta}
\end{equation}
and
\begin{equation}
k_m = \sqrt{\left(\frac{\epsilon + \mu}{\hbar v_F}\right)^2 - (q+M)^2}.
\label{eq:k_m}
\end{equation}
From Eq.~(\ref{eq:k_m}) we see that for a given energy there exists a
critical magnetization
\begin{equation}
M_c = \pm\, 2 |\epsilon + \mu |/(\hbar v_F),
\label{eq:criticalM}
\end{equation}
beyond which for all transverse ($q$) modes the wavefunction changes from
propagating to spatially decaying (evanescent) along the
$x$-direction~\cite{Mondal2010a,Mondal2010b}.
\begin{figure}
\includegraphics[width=.8\columnwidth]{Plots/FNF.eps}
\caption{(Color online) Sketch of the F$_l$NF$_r$
junction. Ferromagnetic films are placed on top of the topological
insulator on the left and right providing exchange fields in these
regions. The arrows indicate the direction of the corresponding
magnetizations $M_l$ and $M_r$, see the text for further
explanation.}
\label{fig:FNFjunction}
\end{figure}
Now we describe the FNF junction, see Fig.~\ref{fig:FNFjunction}.
Region $F_l$ ($ x < 0$) and region $F_r$ ($ x > d $) are modeled as
ferromagnetic regions, respectively, with different magnetizations
$M_l$, $M_r$ along the $y$-axis and corresponding wavefunction
$\psi_F$ [Eq.~(\ref{eq:psiF})]. The Dirac fermions in the middle
region N ($ 0 < x < d$) are described by the wavefunctions $\psi_N$
[Eq.(\ref{eq:eigenstates})]. When calculating transport properties of
the FNF junction, we focus on two different alignments of the
magnetizations of the ferromagnetic regions: the parallel
configuration (M$_l$ $\parallel$ M$_r$), where the magnetizations in
the ferromagnetic regions point in the same direction, and the
anti-parallel configuration (M$_l$ $\parallel$ - M$_r$), in which the
magnetizations are in opposite directions.
Using Eqns.~(\ref{eq:eigenstates})-(\ref{eq:k_m}) we can calculate the
reflection and transmission coefficients for a Dirac fermion with
energy $\epsilon$ and transverse momentum $q$ incident from the left
on the junction, for both the NFN and the FNF junctions. To this end,
we consider a general F$_l$F$_m$F$_r$ junction, where the
wavefunctions in each of the three regions left ($l$), middle ($m$)
and right ($r$) are given by:
\begin{eqnarray}
\psi_{l} &=& \psi_{l}^+ + r_{ll}\, \psi_{l}^- ,
\nonumber \\
\psi_m &=& p\, \psi_m^+ + q\, \psi_m^- ,
\\
\psi_{r} &=& t_{rl}\, \psi_{r}^+.
\nonumber
\end{eqnarray}
Here $\psi_{j}^{\pm}$ ($j = l, m ,r$) are the wavefunctions
(\ref{eq:psiNl}), (\ref{eq:psiNr}) or (\ref{eq:psiF}) (depending on
the junction considered) and $r_{ll}$ and $t_{rl}$ denote the
corresponding reflection and transmission coefficients. By requiring
continuity of the wavefunction at the interfaces $x = 0$ and $x = d$,
we obtain the reflection and transmission coefficients:
\begin{widetext}
\begin{eqnarray}
r_{l l} &=& e^{i \alpha_l}\, \frac{e^{2 i k_m d} (1 + e^{i (\alpha_m+
\alpha_l)}) (e^{i \alpha_m} - e^{i \alpha_r}) + (e^{ i \alpha_l} -
e^{i \alpha_m}) (1 + e^{i (\alpha_m+ \alpha_r)})}{e^{2 i k_m d}
(e^{i \alpha_m} - e^{i \alpha_l}) (e^{i \alpha_m} - e^{i \alpha_r})
+ (1 + e^{i (\alpha_m+ \alpha_l)}) (1 + e^{i (\alpha_m+ \alpha_r)})}
,
\label{eq:generalreflectionLL}\\
t_{rl} &=& \frac{e^{i k_m d}(1 + e^{2 i \alpha_m}) (1 + e^{2 i \alpha_l})} {e^{2
i k_m d} (e^{i \alpha_m} - e^{i \alpha_l}) (e^{i \alpha_m} - e^{i
\alpha_r}) + (1 + e^{i (\alpha_m+ \alpha_l)}) (1 + e^{i (\alpha_m+
\alpha_r)})} .
\label{eq:generaltransmissionLR}
\end{eqnarray}
\end{widetext}
Here $\alpha_j$ denotes the polar angle of the wavevector in region
$j=l,m,r$ [Eqns.~(\ref{eq:alphawithV}) and (\ref{eq:beta})]. When
considering an electron incident from the right lead, one can
similarly obtain $r_{rr}$ and $t_{lr}$.
These expressions for the reflection and transmission coefficients
form the basis of our calculations of the conductance and the pumped
current in Secs.~\ref{sec:conductance} and \ref{sec:pumpedcurrent}
respectively.
\section{Conductance}
\label{sec:conductance}
The conductance $G^{\text{NFN}}$ of a topological insulator NFN
junction has been studied in earlier work by Mondal \textit{et
al.}~\cite{Mondal2010a,Mondal2010b}, who predicted
oscillatory behavior of $G^{\text{NFN}}$ as a function of the applied
bias voltage (see also Fig.~\ref{fig:NFNconductance}). In this section
we first briefly review their results and then add a quantitative
explanation for the oscillations of the conductance. This explanation
is crucial for understanding the behavior of the pumped current in the
next section. We also calculate and analyze the conductance in a FNF
junction.
The general expression for the conductance $G$ across the junction in
terms of the transmission probability $T(\alpha) \equiv |t_{rl}(\alpha)|^2$
is given by
\begin{equation}
G = (G_0 /2) \int_{-\pi/2}^{\pi/2} T(\alpha) \cos \alpha\, d\alpha.
\label{eq:conductance}
\end{equation}
Here $G_0= \frac{2e^2}{h} \rho(eV)\hbar v_F W $, $\rho(eV)= | \mu+eV |
/(2\pi (\hbar v_F)^2)$ denotes the density of states, $W$ is the
sample width, and the integration is over all the angles of incidence
$\alpha$. For $V_l=V_r=0$ ($\alpha_l=\alpha_r=\alpha$) the
angle-dependent transmission probability $T^{\text{NFN}}(\alpha)$ is
given by~\cite{Mondal2010a,Mondal2010b}
\begin{eqnarray}
T^{\text{NFN}}(\alpha) &=&\cos^2(\alpha)\cos^2(\alpha_m)/ \left[
\cos^2(k_md)\cos^2(\alpha) \cos^2(\alpha_m) \right. \nonumber \\
&&\left. + \sin^2(k_md)(1-\sin(\alpha)\sin(\alpha_m))^2 \right],
\label{eq:totaltransmission}
\end{eqnarray}
where $\alpha_m$ is the polar angle of the wave vector in the middle
region as defined in Eq.~(\ref{eq:beta}). This angle can be expressed
in terms of $\alpha$ using the fact that the momentum is conserved
along $y$-axis as:
\begin{equation}
\sin(\alpha_m) = \sin(\alpha) + \frac{M\hbar v_F}{\mid\epsilon+\mu\mid}.
\label{eq:alpham}
\end{equation}
\begin{figure}
\includegraphics[width=.8\columnwidth]{Plots/GNFNvsEn.eps}
\caption{(Color online) The conductance $G^{\text{NFN}}$ of the NFN junction
[Eq.(\ref{eq:conductance})] as a function of $\epsilon / \mu$ for
$V_l=V_r=0$ and for different values of the effective magnetization
$\tilde{M}\equiv \hbar v_F M / \mu = 3$ (solid blue line), 3.5
(dashed green line), 4 (dot-dashed red line) and 4.5 (dotted
light-blue line). The effective junction width $\tilde{d} \equiv \mu
d/(\hbar v_F) = 5$.}
\label{fig:NFNconductance}
\end{figure}
Figure~\ref{fig:NFNconductance} shows the conductance of the NFN
junction [obtained from Eqns.~(\ref{eq:conductance}) and
(\ref{eq:totaltransmission})] as a function of the energy of the
incoming carriers $\epsilon / \mu$ for different values of the
effective magnetization $\tilde{M} \equiv \hbar v_F M / \mu$. For a
given magnetization $M$, the conductance is zero for $\epsilon <
\epsilon_c$, i.e., $G^{\text{NFN}}(\epsilon)=0$, with the critical
energy $\epsilon_c \equiv \hbar v_F M / 2 - \mu$. Below this energy
there are no traveling modes inside the barrier. Our results agree
with the previous results in the
literature~\cite{Mondal2010a,Mondal2010b}.
From Fig.~\ref{fig:NFNconductance} it can be observed that the
conductance changes from plateau-like to oscillatory as $\epsilon/\mu$
increases. In order to provide an explanation for this behavior we
first analyze the plateau like regime in detail. After setting
$\alpha_l = \alpha_r \equiv \alpha$ in
Eq.~(\ref{eq:generalreflectionLL}), we begin by finding the conditions
when the reflection coefficient is zero, i.e., $r_{ll}=0$. The first,
trivial, condition $\alpha = \alpha_m + 2\pi n$ corresponds to the
situation of an entirely normal junction (i.e., no ferromagnetic
region). The second and more interesting condition is $\sin (k_m d) =
0$. This is the case when transmission occurs via a \textit{resonant}
mode of the junction and can be written as (using
Eqns.~(\ref{eq:alpha}) and (\ref{eq:k_m}))
\begin{equation}
k_m d = \frac{|\epsilon + \mu|}{\hbar v_F}d\sqrt{1 - \left( \sin
\alpha + \frac{\hbar v_F M}{|\epsilon + \mu|} \right)^2}=n\pi.
\label{eq:resonantcondition}
\end{equation}
Eq.~(\ref{eq:resonantcondition}) indicates that for a given $M$ and
$\epsilon$ there are certain privileged angles $\alpha_c$ for which the
barrier becomes transparent:
\begin{equation}
\sin (\alpha_c) = \pm \sqrt{1 - \left( \frac{n\pi}{\tilde{d} (1 +
\frac{\epsilon}{\mu})} \right)^2} - \frac{\tilde{M}}{|1 +
\frac{\epsilon}{\mu}|},
\label{eq:condition}
\end{equation}
with $\tilde{d} \equiv d \mu/(\hbar v_F)$ being the dimensionless barrier
length. These modes
are referred to as \textit{resonant} modes in this article.
\begin{figure}%
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn07.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn09.eps}}\\
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn12.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn16.eps}}\\
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn24.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GEn29.eps}}
\caption{(Color online) The transmission probability $T^{\text{NFN}}(\alpha)$
[Eq.~(\ref{eq:totaltransmission})] as a function of the angle of
incidence $\alpha$ for different values of energy $\epsilon/\mu$,
(a) $\epsilon/\mu = 0.7$, (b) $\epsilon/\mu = 0.9$, (c)
$\epsilon/\mu = 1.2$, (d) $\epsilon/\mu = 1.6$, (e)$\epsilon/\mu =
2.4$, and (f) $\epsilon/\mu = 2.9$. Parameters used are $\tilde{d}
= 5$ and $\tilde{M} = 3$.}
\label{fig:Gvsalpha}
\end{figure}
Figure~\ref{fig:Gvsalpha} shows the transmission probability
[Eq.~(\ref{eq:totaltransmission})] as a function of the angle of
incidence $\alpha$ for different values of energy $\epsilon/\mu$. The
dashed (red) vertical lines correspond to the angles satisfying
Eq.~(\ref{eq:condition}) for different $n$. It can be seen that as the
energy increases more resonant modes become available for
transmission. It is also worth noting that for energies at which only
one mode is present ($n=1$, see Fig.~\ref{fig:Gvsalpha}(a)) the
transmission is strongly localized at one particular angle. This
property could be exploited to fabricate single-mode filters. For low
energy excitations only negative angles $\alpha$ (i.e., $q$-momenta
anti-parallel to M) contribute to the conductance, see
Figs.~\ref{fig:Gvsalpha}(a)-(d). As the energy increases, the resonant
modes move from the left to the right and also positive angles
$\alpha$ (i.e., $q$-momenta parallel to M) begin to contribute, see
Figs.~\ref{fig:Gvsalpha}(e) and (f).
\begin{figure}
\includegraphics[width=.8\columnwidth]{Plots/AlphaEn.eps}
\caption{(Color online) The real part of the angle of incidence $\alpha_c$
[Eq.~(\ref{eq:condition})] versus energy $\epsilon/\mu$ for the modes
$n=1,2,3,4,5$. Parameters used are the same as in
Fig.~\ref{fig:Gvsalpha}.}
\label{fig:alphavsen}
\end{figure}
Now we address the question why a mode becomes resonant in the
barrier. Figure~\ref{fig:alphavsen} shows the angle of incidence
$\alpha_c$ [Eq.~(\ref{eq:condition})] for different values of $n$ as a
function of energy $\epsilon/\mu$. For a given $n$, the resonant mode
does not contribute to the conductance if the energy $\epsilon$
satisfies the condition $ \epsilon_c < \epsilon <
\epsilon_c^{\alpha_n}$, because the imaginary part of the momentum
$k_m$ is nonzero and thus the mode is decaying along the
$x$-direction. This critical energy $\epsilon_c^{\alpha_n}$ for each mode
is given by:
\begin{equation}
\frac{\epsilon_c^{\alpha_n}}{\mu}= \frac{n^2
\pi^2}{2\tilde{d}^2 \tilde{M}} + \frac{\tilde{M}}{2} -1.
\label{eq:EnResonantMode}
\end{equation}
When $\epsilon > \epsilon_c^{\alpha_n}$, $k_m$ becomes real
and the mode becomes resonant. As we increase
the energy, all the modes asymptotically reach their saturation angle
$\alpha = \pi/2$.
Finally, we analyze the effect of the appearance of subsequent
resonant modes on the conductance.
For small energies, the conductance increases in plateau-like steps
(see Fig.~\ref{fig:NFNconductance}). The first plateau corresponds to
the situation in which the first transmission mode appears at
$\alpha=-\pi/2$. As the energy increases, a new resonant mode appears
and the conductance increases in a step-like manner. The plateaus are
not sharp due to the fact that each new mode appearing is not sharply
peaked, but rather has a certain distribution around a particular
angle of incidence, see Fig.~\ref{fig:Gvsalpha}. Once the energy is
large enough for there to be contributions from both positive and
negative angles of incidence, the conductance becomes oscillatory. For
very large energies ($\epsilon \gg \epsilon_c$), the effect of the
magnetic barrier disappears and the conductance becomes unity
($G^{\text{NFN}}= G_0$).
\begin{figure}[tb]
\includegraphics[width=.8\columnwidth]{Plots/GNFNvsEnV.eps}
\caption{(Color online) The conductance $G^{\text{NFN}}$ of the NFN
junction as a function of $\epsilon / \mu$ for different values of
gate voltages, $eV/\mu=0$ (solid blue line), 1 (dashed green line),
2 (dashed-dot red line) and 3 (double-dotted light-blue line). As
before, $\tilde{M} = 3$ and $\tilde{d} =5$.}
\label{fig:N1MN2conductanceOverV}
\end{figure}
Figure~\ref{fig:N1MN2conductanceOverV} shows the conductance as a
function of energy for several values of applied bias voltages $V_l =
V_r \equiv V$. As expected, the features of the conductance remain the
same for finite $V$. As we increase $eV$, the critical energy
$\epsilon_c = \hbar v_F M/2 +eV/2 -\mu$ for the onset of the
conductance increases and the spacing between two consecutive resonant
modes decreases. As a result the plateaus become narrower.
In the remaining part of this section we study the conductance in a
topological insulator FNF junction, as shown in
Fig.~\ref{fig:FNFjunction}. We consider both the junction with
parallel and with anti-parallel magnetization in the ferromagnetic
regions. In the parallel configuration, using $\alpha_l = \alpha_r
\equiv \alpha = \sin^{-1} ( \hbar v_F (q+M)/ |\epsilon + \mu| )$ and
$\alpha_{m} = \sin^{-1} ( \hbar v_F q/|\epsilon + \mu|)$ in
Eq.~(\ref{eq:generaltransmissionLR}), we find that the conductance is
similar to the conductance of a NFN junction, as displayed in
Fig.~\ref{fig:NFNconductance}. However, in the FNF junction the first
resonant mode becomes resonant for positive $\alpha$ (i.e., transverse
$q$-momentum parallel to $M$) and as the energy increases, the
resonances move towards negative values of the angle $\alpha$. This,
however, does not affect the total conductance, as we sum over all
possible angles of incidence, and the same analysis as for the NFN
junction presented above can be applied to understand the FNF junction
with parallel magnetization.
In the case of anti-parallel alignment of the magnetization in the two
ferromagnetic regions, we substitute $\sin(\alpha_l) = \hbar v_F (q+M)/
|\epsilon + \mu|$, $\sin(\alpha_r) = \hbar v_F (q-M)/ |\epsilon + \mu|$
and $\sin(\alpha_{m}) =\hbar v_F q/|\epsilon + \mu|$ in
Eq.~(\ref{eq:generaltransmissionLR}). The conductance of this junction
was studied previously in Refs.~\cite{Wu2010,Salehi2011} and the
transmission probability $T^{\text{FNF,AP}}(\alpha_l,\alpha_r) \equiv
|t_{rl}(\alpha_l,\alpha_r)|^2 $ is given by:
\begin{widetext}
\begin{equation}
T^{\text{FNF,AP}}(\alpha_l,\alpha_r)=
\frac{\cos^2(\alpha_l) \cos^2(\alpha_m)}
{\cos^2(k_md) \cos^2(\frac{\alpha_l + \alpha_r}{2}) \cos^2(\alpha_m) +
\sin^2(k_md) \left[ \cos \left( \frac{\alpha_l - \alpha_r}{2} \right) -
\sin(\frac{\alpha_l + \alpha_r}{2}) \sin(\alpha_m) \right]^2}.
\label{eq:antiparallel}
\end{equation}
\end{widetext}
The total conductance is obtained by multiplying
$T^{\text{FNF,AP}}(\alpha_l,\alpha_r)$ with $\cos(\alpha_r)/\cos(\alpha_l)$
and then integrating over the allowed angles of
incidence~\cite{restriction}, i.e., from $\alpha_{c_1} = \sin^{-1} (2
\hbar v_F M/(|\epsilon + \mu|) - 1)$ to $\alpha_{c_2} =\sin^{-1} (2
\hbar v_F M/(|\epsilon + \mu|) +1)$. Thus we can write
\begin{equation}
G^{\text{FNF,AP}}=G_0/2\int_{\alpha_{c_1}}^{\pi/2}G^{\text{FNF,AP}}
(\alpha_l,\alpha_r)\cos(\alpha_l)d\alpha_l,
\label{eq:newequation}
\end{equation}
where
\begin{equation}
G^{\text{FNF,AP}}(\alpha_l,\alpha_r) = \frac{\cos(\alpha_r)}{\cos(\alpha_l)}
T^{\text{FNF,AP}}(\alpha_l,\alpha_r).
\label{eq:neweq2}
\end{equation}
\begin{figure}
\includegraphics[width=.8\columnwidth]{Plots/GFNFanticonf.eps}
\caption{(Color online) The conductance $G^{\text{FNF,AP}}$ of the FNF junction in
the anti-parallel configuration as a function of $\epsilon /\mu$,
for $\tilde{M} = 3$ (blue solid line), $3.5$ (green dashed line) and
$4$ (red dotted line). The parameter $\tilde{d}=5$.}
\label{fig:FNFantiparallelVSparallel}
\end{figure}
Fig.~\ref{fig:FNFantiparallelVSparallel} shows the conductance of the
FNF junction in the anti-parallel configuration. From the horizontal
axis we see that the critical energy $\epsilon_c$ for the onset of the
conductance is larger than in the corresponding parallel
configuration. Moreover, as the energy $\epsilon/\mu$ increases the
conductance exhibits no plateau behavior: it increases in an
oscillatory fashion. This oscillatory behavior can be understood from
Fig.~\ref{fig:GANTIvsalpha}, which shows $G^{\text{FNF,AP}}(\alpha)$
for four different values of $\epsilon/\mu$. Note that all the angles
($\alpha_l$, $\alpha_m$ and $\alpha_r$) can be expressed in terms of
one angle, which we choose to be $\alpha_l\equiv\alpha$. As the energy
increases, the area under the curve oscillates resulting in
oscillations in the conductance.
Summarizing, we have obtained a quantitative explanation for the
behavior of the conductance in topological insulator NFN and
FNF-junctions in terms of the number of resonant modes in the
junction. This explanation forms the basis for understanding the
behavior of the pumped current in the next section.
\begin{figure}
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GANTI247.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GANTI257.eps}}
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GANTI310.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GANTI340.eps}}
\caption{The angle-dependent total transmission
$T^{\text{FNF,AP}}(\alpha_l,\alpha_r)$ for the anti-parallel
configuration of the FNF junction as a function of the angle of
incidence~$\alpha_l$ for (a) $\epsilon/\mu = 2.47$, (b)
$\epsilon/\mu = 2.57$, (c) $\epsilon/\mu = 3.10$ and (d)
$\epsilon/\mu = 3.40$. Parameters used are $\tilde{M} = 3$ and
$\tilde{d} = 5$.}
\label{fig:GANTIvsalpha}
\end{figure}
\\
\section{Adiabatically pumped current}
\label{sec:pumpedcurrent}
In this section we investigate adiabatically pumped currents through
NFN and FNF junctions in a topological insulator~\cite{difference}.
In general, a pumped current is generated by slow variation of two
system parameters $X_1$ and $X_2$ in the absence of a bias
voltage~\cite{Buttiker1994,Brouwer1998}. For periodic modulations
$X_1(t)= X_{1,0} + \delta X_1 \cos (\omega t)$ and $X_2(t) = X_{2,0} +
\delta X_2 \cos (\omega t + \phi)$, the pumped current $I_p$ into the
left lead of the junction can be expressed in terms of the area $A$
enclosed by the contour that is traced out in $(X_1,X_2)$-parameter
space during one pumping cycle as~\cite{Brouwer1998}:
\begin{subequations}
\begin{eqnarray}
I_p & = & \frac{\omega e}{2 \pi^2} \int_A\, dX_1\, dX_2\,
\sum_{m} \ \Pi (X_1,X_2)
\label{eq:Ip} \\
& \approx & \frac{\omega e}{2 \pi}\, \delta X_1\, \delta X_2\, \sin
\phi \, \sum_{m} \ \Pi (X_1,X_2),
\label{eq:IpLinearResp}
\end{eqnarray}
\end{subequations}
with
\begin{equation}
\Pi (X_1,X_2) \equiv \mbox{\rm Im}\ \left( \frac{\partial
r_{ll}^{\ast}}{\partial X_1} \frac{\partial r_{ll} }{\partial X_2} +
\frac{\partial t_{lr}^{\ast}}{\partial X_1} \frac{\partial
t_{lr}}{\partial X_2} \right).
\label{eq:Pifunc}
\end{equation}
Here $r_{ll}$ and $t_{lr}$ represent the reflection and transmission
coefficients into the left lead and the index $m$ sums over all modes (a
similar expression can be obtained for the pumped current into the right
lead). Eq.~(\ref{eq:IpLinearResp}) is valid in the bilinear response
regime where $\delta X_1 \ll X_{1,0}$ and $\delta X_2 \ll X_{2,0}$ and
the integral in Eq.~(\ref{eq:Ip}) becomes independent of the pumping
contour.
First we analyze the NFN pump, where the pumped current is generated
by adiabatic variation of gate voltages $V_l$ and $V_r$ which change
the chemical potential in the normal leads on the left and right of
the junction, respectively (see
Fig.~\ref{fig:NFNjunction}). Calculating the derivatives of the
reflection and transmission coefficients $r_{ll}$
[Eq.~(\ref{eq:generalreflectionLL})] and $t_{rl}$ with respect to
$\alpha_l$ and $\alpha_r$, substituting into Eq.~(\ref{eq:Pifunc}) and
using $\partial \alpha_j /(e\partial V_j) = \tan (\alpha_j)/ |\epsilon
+ \mu -eV_j|$ ($j = l, r$), the pumped current for $V_1 = V_2 \equiv
V$ and for a specific angle of incidence $\alpha$ is given by:
\begin{widetext}
\begin{equation}
I_{p}^{\text{NFN}}(\alpha)= - I_0^{\text{NFN}}
\frac{\cos^3(\alpha_m) \sin^2 (\alpha) \cos (\alpha) \sin( 2 k_m d) } {(1 +
\epsilon/ \mu- eV/ \mu)^2 (\cos^2 (\alpha) \cos^2(\alpha_m) \cos^2
(k_m d) +\sin^2 (k_m d) (1 - \sin (\alpha) \sin (\alpha_m))^2)^2}.
\label{eq:NFNpumpedcurrent}
\end{equation}
\end{widetext}
Here $I_0^{\text{NFN}} \equiv \omega e/(8\pi) \sin({\phi}) (e \delta
V_1 /\mu)(e\delta V_2/\mu)$ and
$\sin(\alpha_m)$ is given by Eq.~(\ref{eq:alpham}). In the limit $M
\rightarrow 0$ (i.e., $\alpha_m \rightarrow \alpha$) in an entirely normal
junction, we obtain from Eq.~(\ref{eq:NFNpumpedcurrent}) the
angle-dependent pumped current as:
\begin{equation}
I_p^{\text{NFN}}|_{\tilde{M} =0}= I_0^{\text{NFN}}
\int_{-\pi/2}^{\pi/2} \frac{\sin (2 k_m d) \sin^2 \alpha }{ (1 +
\epsilon /\mu -eV /\mu ) \cos^3 \alpha} \, d\alpha.
\label{eq:NFNtotalcurrentm=0}
\end{equation}
On the other hand, the transmission $T^{\text{NFN}}(\alpha)$
[Eq.~(\ref{eq:totaltransmission})] in this limit is given by
$T^{\text{NFN}}|_{\tilde{M} \rightarrow 0} \rightarrow 1$, independent
of the angle of incidence $\alpha$. We notice that even if the
probability for transmission is one, it is possible to pump a current
in the adiabatic driving regime. The total pumped current $I_p^{NFN}$
is then obtained by integrating over $\alpha$:
\begin{equation}
I_p^{\text{NFN}}= \int_{-\pi/2}^{\pi/2} I_p^{\text{NFN}} (\alpha)
\cos\alpha\, d\alpha.
\label{eq:NFNtotalcurrent}
\end{equation}
In general, this integral cannot be evaluated analytically and we have
obtained our results numerically.
\begin{figure}
\centering
\subfigure{\includegraphics[width=.8\columnwidth]{Plots/IpNFNvsEn.eps}}
\\
\caption{The pumped current $I_p^{\text{NFN}}$
[Eq.~(\ref{eq:NFNtotalcurrent})] in the NFN junction as a function
of $\epsilon/\mu$ for $V =0$, $\tilde{d} = 5$ and $\tilde{M}= 3$. }
\label{fig:NFNIp}
\end{figure}
Figure~\ref{fig:NFNIp} shows the total pumped current
$I_p^{\text{NFN}}$ (in units of $I_0^{\text{NFN}}$) at zero bias $V =
V_1=V_2=0$ for $\tilde{M}=3$.
Comparing Figs.~\ref{fig:NFNconductance} and~\ref{fig:NFNIp} we see
that there is a correlation between the pumped current and the
conductance for the NFN junction: for low energies, the pumped current
$I_p^{NFN}$ is zero as no traveling modes are allowed in the
junction. As we increase the energy, each time a resonant mode appears
[see Eq.~(\ref{eq:resonantcondition})], the pumped current diverges
and changes sign. For energies where both positive and negative angles
of incidence contribute to the conductance, the pumped current remains
finite but keeps changing its sign. From
Figs.~\ref{fig:NFNconductance} and~\ref{fig:NFNIp} it can also be seen
that the pumped current vanishes for energies at which subsequent
resonant modes become fully transmitting. In order to gain further
insight we plot the analogue of Fig.~\ref{fig:Gvsalpha} for the pumped
current. Figure ~\ref{fig:IpNFNvsalpha} shows the pumped current
[Eq.~\ref{eq:NFNtotalcurrent}] as a function of the angle of incidence
$\alpha$ for different values of $\epsilon/\mu$. The chosen values of
$\epsilon/\mu$ are same as in Fig.~\ref{fig:Gvsalpha}. We see that the
features in Fig.~\ref{fig:IpNFNvsalpha} have a direct correlation with
the features in Fig.~\ref{fig:Gvsalpha}: whenever there is a sharp
peak in the transmission the pumped current diverges and changes
sign. The key feature that distinguishes between the pumped current
and the conductance is that the pumped current changes sign at
particular values of the energy, while the conductance does not.
Now we analyze the pumped current in the FNF junction with parallel
orientation of the magnetizations. In this system, the driving
parameters are the magnetizations $M_l$ and $M_r$ in the left and
right contacts, respectively, see Fig.~\ref{fig:FNFjunction}. After
calculating the derivatives of the reflection and transmission
coefficients, obtaining the imaginary part of Eq.~(\ref{eq:Pifunc}),
and using $\partial \alpha_j/\partial M_j = \hbar v_F /( |\epsilon +
\mu|\cos(\alpha_j) )$ ($j=l,r)$, the pumped current
$I_p^{\text{FNF}}(\alpha)$ for $M_1=M_2=M$ is:
\begin{widetext}
\begin{equation}
I_p^{\text{FNF}}(\alpha)= - I_0^{\text{FNF}}
\frac{\cos^3(\alpha_m) \cos(\alpha)\sin (2 k_m d)}{(1 +
\epsilon/\mu)^2 (\cos^2 (\alpha) \cos^2 (\alpha_m) \cos^2 (k_m d)
+\sin^2 (k_m d) (1 - \sin (\alpha) \sin (\alpha_m))^2)^2}.
\label{eq:FNFpumpedcurrent}
\end{equation}
\end{widetext}
Here $I_0^{\text{FNF}}= \omega e/(8\pi) \sin ({\phi}) (\hbar v_F\delta
M_l/\mu)(\hbar v_F \delta M_r/\mu)$ and $\sin (\alpha_m)= \sin
(\alpha) -\hbar v_F M/(|\epsilon + \mu|)$.
\begin{figure}
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha07.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha09.eps}}\\ \subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha12.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha16.eps}}\\ \subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha24.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsalpha29.eps}}
\caption{The pumped current $I_{p}^{\text{NFN}}(\alpha)$
[Eq.~(\ref{eq:NFNpumpedcurrent})] as a function of the angle of
incidence $\alpha$ for different values of energy $\epsilon/\mu$,
(a) $\epsilon/\mu = 0.7$, (b) $\epsilon/\mu = 0.9$, (c)
$\epsilon/\mu = 1.2$, (d) $\epsilon/\mu = 1.6$, (e) $\epsilon/\mu =
2.4$, and (f) $\epsilon/\mu = 2.9$. Parameters used are $\tilde{d}
= 5$ and $\tilde{M} = 3$. }
\label{fig:IpNFNvsalpha}
\end{figure}
\begin{figure}
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/GNFNvsD.eps}}\hfill
\subfigure{\includegraphics[width=.48\columnwidth]{Plots/IpNFNvsD.eps}}
\caption{ (a) The conductance of the NFN junction as a function of
$\tilde{d}$. (b) The pumped current for the NFN junction as a
function of $\tilde{d}$. Parameters used are $\epsilon/\mu=2.5$,
$\tilde{M} = 3$ and $V =0$. }
\label{fig:IpNFNvsD}
\end{figure}
The behavior of the current $I_p^{\text{FNF}}$ is similar to that of
the pumped current $I_p^{\text{NFN}}$ in a NFN-junction (shown in
Fig.~\ref{fig:NFNIp}). This can also be seen by comparing the
denominators in Eqns.~(\ref{eq:NFNpumpedcurrent}) and
(\ref{eq:FNFpumpedcurrent}). Again we observe that the pumped current
diverges at exactly the same locations where the conductance changes
sharply. But there is an important difference between both pumped
currents. The pumped current in an NFN-junction at normal incidence
vanishes, $I_p^{\text{NFN}}(\alpha=0) =0$, while
$I_p^{\text{FNF}}(\alpha=0)\ne 0$.
This difference arises because the two pumps are driven by two
different parameters (voltages in the NFN pump and magnetizations in
the FNF pump).
Finally, we briefly analyze the behavior of the pumped current as a
function of the width $d$ of the middle region. For energies below
$\epsilon_c$, the pumped current of the NFN junction decays to zero as
the width $d$ increases (there are no resonant modes in the
system). For energies larger than $\epsilon > \epsilon_c$, the pumped
current $I_p^{\text{NFN}}$ oscillates as a function of width
$\tilde{d}$. Fig.~\ref{fig:IpNFNvsD} shows the conductance and the
pumped current as a function of $\tilde{d}$ for
$\epsilon/\mu=2.5$. The peaks in the conductance correspond to the
resonance condition Eq.~(\ref{eq:resonantcondition}). The pumped
current $I_p^{\text{NFN}}$ changes its sign at exactly the same values
of $\tilde{d}$ where the conductance has a maximum.
This analysis holds as well for the FNF junction.
\section{Summary and discussion}
\label{sec:summary}
To summarize, we have analyzed quantum transport by Dirac fermion
surface states in NFN and the FNF junctions in a 3D topological
insulator. We have shown that for low energies the appearance of a new
resonant mode results in a plateau-like increment of the conductance
and a diverging pumped current in these junctions which also changes
sign. This is our key result, and represents an experimentally
distinguishable signature between conductance and the pumped current.
We highlighted an interesting difference between the two different
pumping mechanisms for the NFN and FNF junctions, observing different
behaviors for normal incidence ($\alpha=0$). Experimentally, the NFN
pump could be realized using current technology. The FNF pump will be
more difficult to realize since it requires oscillating
magnetizations. A possible way to realize a FNF pump could be by
moving the two ferromagnetic layers coherently using a nanomechanical
oscillator~\cite{Kovalev2005}. Experimental verification of our
predictions will provide further insight into quantum transport
through these junctions.
\begin{acknowledgments}
This research was supported by the Dutch Science Foundation NWO/FOM.
\end{acknowledgments}
|
\section{Introduction}
We provide representations of the scaling functions
of the three-dimensional $O(4)$ model which can be used in tests of other
models on their membership of the corresponding universality class.
Contrary to the often analyzed scaling function of the order parameter,
the so-called magnetic equation of state, our main interest here is to
determine directly the scaling function of the free energy density and
its derivatives. This
is especially of importance for applications to quantum chromodynamics (QCD)
with two degenerate light-quark flavors at finite temperature.
Two-flavor QCD is believed \cite{Pisarski:ms}-\cite{Ejiri:2009ac} to
belong to the 3d $O(4)$ universality class at its chiral transition in the
continuum limit. In the vicinity of the chiral phase transition temperature
the reduced temperature variable in QCD also depends quadratically on
the quark chemical potential. Derivatives of the singular part of the
free energy density of QCD with respect to chemical potential, which define
cumulants of fluctuations of net quark number, thus are controlled by
scaling functions that are given by derivatives of the scaling
function of the free energy density in a three-dimensional $O(4)$ model.
Obtaining explicit parametrizations of higher order derivatives of the
scaling functions of the free energy density became of interest recently as these
higher order derivatives control the scaling behavior of fluctuations
of conserved charges, e.g. the net baryon number \cite{redlich}.
These quantities are currently measured at RHIC \cite{Star} and will also
be measured in heavy ion experiments at the LHC.
\section{\boldmath The three dimensional $O(4)$ model}
The specific model which we study here is the standard $O(4)$-invariant
nonlinear $\sigma$-model, which is defined by
\begin{equation}
\beta\,{\cal H}\;=\;-J \,\sum_{<{\vec x},{\vec y}>}\vec \phi_{\vec x}\cdot
\vec \phi_{\vec y} \;-\; {\vec H}\cdot\,\sum_{{\vec x}} \vec \phi_{\vec x} \;,
\end{equation}
where ${\vec x}$ and ${\vec y}$ are nearest-neighbor sites on a
three-dimensional hypercubic lattice, and $\vec \phi_{\vec x}$ is a
four-component unit vector at site ${\vec x}$. The coupling $J$ and
the external magnetic field $\vec H$ are reduced quantities, that is
they contain already a factor $\beta=1/T$. In fact, we consider in
the following the coupling directly as the inverse temperature,
$J\equiv 1/T$.
The partition function is then
\begin{equation}
Z(T,H)\;=\; \int \prod_{\vec x} d^{\,4}\phi_{\vec x}\;\delta (\vec \phi_{\vec x}^{\,2}
-1) \exp(-\beta\, {\cal H} ) ~.
\end{equation}
We introduce the order parameter $M$ as the derivative of the
free energy density, $f(T,H) \;=\; -\frac{1}{V}\ln Z$, with respect
to the magnitude of the external magnetic field $\vec H = H\vec e_H$,
\begin{equation}
M \;=\; - \frac{\partial f}{\partial H}\; =\;\<\, \phi^{\parallel} \,\>~,
\label{Mterm}
\end{equation}
where $\phi^{\parallel}$ is the field component parallel to the magnetic
field $\vec H$.
In the vicinity of the critical point the free energy density may be splitted
into a singular (non-analytic) ($f_s$) and a non-singular ($f_{ns}$) part,
\begin{equation}
f(T,H)\;=\; f_s(T,H) + f_{ns}(T)~.
\end{equation}
The singular part is a homogeneous function of the variable $h=H/H_0$
and the reduced temperature $\bar{t}=(T-T_c)/T_0$, where $H_0$ and $T_0$
set the scale in the critical region. The singular part may be expressed
in terms
of a universal scaling function $f_f$, which itself only depends on the scaling
variable $z= \bar t/h^{1/\Delta}$, i.e.,
\begin{equation}
f_s \;=\; H_0 h^{1+1/\delta} f_f(z)~.
\label{fsscale}
\end{equation}
Here we have introduced the gap exponent, $\Delta=\beta\delta$, which is
given in terms of the more commonly used critical exponents $\beta$ and
$\delta$. The latter define the scaling properties of the order
parameter as function of temperature at $h=0$ and
as function of the external field at $t=0$, respectively.
Eq.~\ref{fsscale} establishes the relation between the universal scaling
function of the order parameter ($f_G$) and the scaling function of the free
energy density ($f_f$). Using Eq.~\ref{Mterm} we find
\begin{eqnarray}
M &=& h^{1/\delta} f_G(z) \\
f_G(z) &=& -\left(1+\frac{1}{\delta}\right) f_f(z)
+\frac{z}{\Delta}f_f^\prime(z)~.
\label{fgdiff}
\end{eqnarray}
In the following we will exploit the differential equation, Eq.~\ref{fgdiff},
to determine the scaling function $f_f(z)$ from $f_G(z)$.
\section{Scaling functions of the free energy density and the order parameter}
As the universal scaling functions of the free energy density, $f_f(z)$, and
the order parameter, $f_G(z)$, are related through the differential equation,
Eq.~\ref{fgdiff}, the knowledge of $f_G(z)$ is sufficient to determine $f_f(z)$.
We summarize in the following the relevant relations that determine $f_f(z)$,
once a suitable parametrization of $f_G(z)$ is known. Further details are
given in Ref.~\cite{O4}.
We consider a parametrization of $f_G(z)$ by introducing three series
expansions that are valid for small $z$ and in the asymptotic regions
$z\rightarrow \pm \infty$, respectively,
\begin{equation}
f_G(z) = \begin{cases}
\sum_{n=0}^\infty b_nz^n & ,\; z\;\; \mbox{small} \\
z^{-\gamma} \cdot \sum_{n=0}^\infty d_n^+ z^{-2n\Delta} &\; ,\; z\rightarrow +\infty \\
(-z)^{\beta} \cdot \sum_{n=0}^\infty d_n^- (-z)^{-n\Delta/2} &\; ,\; z\rightarrow -\infty
\end{cases}
\label{fGparam}
\end{equation}
The corresponding parametrization for the scaling function of the
free energy density is then given by,
\begin{equation}
f_f(z) = \begin{cases}
\sum_{n=0}^\infty a_nz^n & ,\; z\;\; \mbox{small} \\
z^{2-\alpha} \cdot \sum_{n=0}^\infty c_n^+ z^{-2n\Delta} &\; ,\;
z\rightarrow +\infty \\
(-z)^{2-\alpha} \cdot \sum_{n=0}^\infty c_n^- (-z)^{-n\Delta/2} &\; ,\;
z\rightarrow -\infty
\end{cases}
\label{ffparam}
\end{equation}
The relation between the expansion coefficients in the series representations
for $f_G$ and $f_f$ are easily obtained by using the differential equation,
Eq.~\ref{fgdiff}, and comparing coefficients for $n\ge 0$,
\begin{equation}
a_n\;=\;\frac{\Delta b_n}{\alpha+n-2} \;\; ,\; \;
c_{n+1}^+\;=\; \frac{-d_n^+}{2(n+1)} \;\;, \; \;
c_{n+2}^- \;=\; -\frac{2d_n^-}{n+2} \;\;,\;
\label{coeff}
\end{equation}
and $c_1^-= 0$. This leaves the coefficients $c_0^\pm$ still undetermined.
They can be obtained as
\begin{eqnarray}
c_0^+ &=& \frac{\Delta}{2-\alpha} \int_0^\infty dy\ y^{\alpha-2}
\left[ f_G^\prime(y)-f_G^\prime(0)-yf_G''(0) \right]~, \\
c_0^- &=& \frac{-\Delta}{2-\alpha} \int^0_{-\infty} dy\ (-y)^{\alpha-2}
\left[ f_G^\prime(y)-f_G^\prime(0)-yf_G''(0) \right]~.
\label{c_0}
\end{eqnarray}
Here $\alpha$ is the specific heat critical exponent, which is negative
in the three dimensional $O(4)$ universality class.
\begin{table}[t]
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
$b_0$ & $b_1$ &$b_2$ & $b_3$ \\
\hline
$1$ & $-0.3166125\pm 0.000534$ & $-0.04112553\pm 0.001290$ & $0.00384019\pm 0.000667 $ \\
\hline
~ & $b_4^+$ &$b_5^+$ & $b_6^+$ \\
\hline
~& $0.006705475\pm 0.001704$ & $0.0047342\pm 0.001429$ & $-0.001931267\pm 0.000312$\\
\hline
~ & $b_4^-$ &$b_5^-$ & $b_6^-$ \\
\hline
~ & $0.007100450\pm 0.000160$ & $0.0023729\pm 0.000095$ & $0.000272312\pm 0.000021$ \\
\hline
\end{tabular}
\end{center}
\caption{\label{tab:expansion_b}Coefficients of the small $z$-expansion
of the scaling function $f_G(z)$ of the order parameter. For $n\ge 4$ we
give different expansion coefficients for negative and positive $z$-values.}
\end{table}
The coefficients $a_n$ for all $n$ and $c_n$, for $n >0$, are obtained from
a parametrization of the scaling function of the order parameter. The
relevant expansion coefficients $b_n$, $b_n^\pm$ and $d_n^\pm$ have been
obtained from numerical data for the order parameter itself as well
as its susceptibility \cite{O4}. In this step one explicitly makes
use of a set of values for critical exponents in the three dimensional
O(4) universality class. We used: $\beta=0.380$ and~ $\delta=4.824$.
All other critical exponents can be derived using hyperscaling relations.
E.g., the specific heat exponent is given by, $\alpha=-0.213$.
We list the resulting expansion coefficients
in Table~\ref{tab:expansion_b} and \ref{tab:expansion_d}. Note that we
give different expansion coefficients $b_n^+$ and $b_n^-$ for $n\ge 4$
to better reproduce the asymmetric form of the scaling function $f_G(z)$
also for small values of $z$ with only a small number of expansion
coefficients. The corresponding scaling function of the order parameter
and its first derivative is shown in Fig.~\ref{fig:fG}.
Having at hand a parametrization of $f_G(z)$ we finally can determine
the remaining coefficients $c_0^\pm$, which complete the parametrization
of $f_f(z)$. These expansion coefficients are listed in
Table~\ref{tab:expansion_c}.
\begin{figure}[htp]
\centerline{
\epsfig{file=fig1pos, width = 0.5\textwidth}
\epsfig{file=fig2pos, width = 0.5\textwidth}
}
\caption{\label{fig:fG} The scaling function of the order parameter
(left) and its derivative (right).}
\end{figure}
\begin{figure}[htp]
\centerline{
\epsfig{file=ff0, width = 0.45\textwidth}
\epsfig{file=ff1_2, width = 0.45\textwidth}
}
\centerline{
\epsfig{file=ff2_2, width = 0.45\textwidth}
\epsfig{file=ff3c, width = 0.45\textwidth}
}
\caption{\label{fig:ff} The scaling function of the free energy density and
its first three derivatives.}
\end{figure}
\begin{table}[ht]
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$d_0^+$ & $d_1^+$ &$d_2^+$ \\
\hline
$1.10599\pm 0.00555$ & $-1.31829\pm 0.1087$ & $1.5884\pm 0.4646$ \\
\hline
\hline
$d_0^-$ & $d_1^-$ &$d_2^-$ \\
\hline
$1$ & $0.273651\pm 0.002933$& $0.0036058\pm 0.004875$\\
\hline
\end{tabular}
\end{center}
\caption{\label{tab:expansion_d}Coefficients for the asymptotic series
expansions of $f_G(z)$ in the region of large positive and negative $z$-values,
respectively.}
\end{table}
\begin{table}[t]
\begin{center}
\begin{tabular}{|c|c|}
\hline
$c_0^+$ & $c_0^-$ \\
\hline
$0.422059886\pm 0.010595$ & $0.229176194\pm 0.010669$ \\
\hline
\end{tabular}
\end{center}
\caption{\label{tab:expansion_c}The leading expansion coefficients
for the singular part of the free energy density.}
\end{table}
\section{Discussion and Conclusions}
The availability of high accuracy numerical data on the order parameter
and its susceptibility in a three dimensional, $O(4)$ symmetric spin
model allowed us to extract the underlying scaling function of the
free energy density and its first three derivatives. As the specific heat
exponent $\alpha$ is negative in the 3d $O(4)$ universality class,
it is only the third
derivative with respect to temperature, which diverges at the critical
point. The corresponding scaling function $f'''_f(z)$ has two extrema;
a rather shallow minimum in the symmetry broken phase and a pronounced
maximum in the symmetric phase. The latter is located at
$z_p^{3,0}\simeq 1.45$. This happens to be close to the location of
the peak in the susceptibility of the order parameter,
$z_p^{0,2}=1.374(3)$.
The higher order derivatives of the scaling function of the free
energy density play a central role in the discussion of fluctuations of
conserved charges in QCD, e.g. the singular behavior of the $2n$-th
order cumulant of net baryon number fluctuations is related to the
$n$-th derivative of $f_f(z)$. The change of sign of $f'''_f(z)$ and
its pronounced maximum characterize the QCD transition. In fact, the
change of sign of $f'''_f(z)$ suggests that $6$th order cumulants of
net baryon number are negative in the vicinity of the QCD transition
line. This may be detectable in a heavy ion collision, if the production
of hadrons (freeze-out) occurs at temperatures and baryon chemical
potentials that are close to the QCD crossover transition line.
|
\section{SNR W44}
{The middle-aged (~20000 yr) supernova remnants W44 is located at 3.1 kpc from us.}
The AGILE/GRID instrument detected gamma-ray emission from SNR W44 in the energy range 50 MeV - 10 GeV with a significance of 15.8 sigma;
gamma-ray ditribution shows an extended source with a morphology well correlated with the radio shell \cite{giuliani11}.
\begin{figure}[t!]
\begin{center}
\resizebox{5.5cm}{!}{\includegraphics[clip=true]{w44.jpg}}
\end{center}
\caption{\footnotesize
SNR W44 as seen by AGILE for energies greater than 400 MeV
}
\label{w44}
\end{figure}
The AGILE energy band is complementary with respect to the band 0.2-30 GeV already investigated by the Fermi/LAT instrument for the same
SNR \citep{abdo10}.
The combination of the AGILE/GRID and Fermi/LAT data allows to obtain a spectrum showing, with unprecedented precision, that the gamma-ray emission from W44 is described by a broad peak around 1 GeV (see figure \ref{hadronic_model}).
With such an accurate spectrum, we can precisely deduce the spectral parameters of the parent particle population (energy of the spectral break, spectral index below and above the break) and dimonstrate that these are not compatible with the electron population seen by radio continuum observations \citep{castelletti07}.\\
An hadronic scenario can instead adequately explain this feature in the gamma-ray spectrum.
In figure \ref{hadronic_model}, AGILE spectral data, together the radio spectral data, are fitted with an hadronic model
characterized by a magnetic field B=70 $\mu$G and a density n =100 cm$^{-3}$.
A fit of the multi-wavelength data set with leptonic models but no good combination of the parameters exists that can fit W44 spectrum.
\begin{figure}[h!]
\begin{center}
\includegraphics[bb=0 0 383 246, scale=0.6]{hadronic_model_b70.jpg}
%
\end{center}
\caption{Hadronic model, characterized by B
=70 µG and n =100 cm−3, of the broad-band spectrum of
SNR W44 superimposed with the radio (data points in red color)
and gamma-ray data of Fig. 2 (in blue color). TeV upper limits
are also shown. The yellow curve shows the neutral
pion emission from the accelerated proton distribution. The green curves
show the electron contribution by
synchrotron (dashed curve), Bremsstrahlung (solid curve), and IC
(dotted curve) emissions. The red curve shows the total gamma-
ray emission.}
\label{hadronic_model}
\end{figure}
AGILE for the first time proved that Galactic Cosmic Rays are accelerated by SNRs.
Moreover, W44 spatial distribution shows that the gamma-ray emission originates in the shell of the SNR.
The large dimensions of the source, combined with the good imaging capabilities of AGILE for E$>$400 MeV, allow us to detect spectral variation
along the SNR shell.
The energy-resolved spatial distribution shows that the bulk of the emission in the 0.4-1 GeV band is generated in the northern part of
the shell while most of the emission in the 1-3 GeV energy band comes from the southern part of the shell (see figure \ref{color}).
These observations are in agreement with several theoretical models which predict that, at any given time, protons are monochromatically
injected in the ISM with an energy depending on the magnetic field at that time (see, for example, \citealt{gabici09}).
This will allow us to deduce the energy of the protons currently emitted by the different part of the SNR, which is crucial for understanding
the overall acceleration mechanism in SNRs.
\begin{figure}[h!]
\resizebox{\hsize}{!}{\includegraphics[clip=true]{color2.jpg}}
\caption{
\footnotesize
AGILE intensity map for the W44 region in the energy range 400MeV-1GeV (Red) and 1-3 GeV (blue).
Green contours show the radio continuum flux density at 1,4 GHz as observed by the VLA (Castelletti et al, 2007)
}
\label{color}
\end{figure}
\section{SNR W28}
W28 is a mixed morphology SNR with an age of more than 35 000 years located at a distance of about 1.9 kpc.
A system of massive molecular clouds is associated to the SNR as revealed by the CO (J = 0 $\rightarrow$ 1) observation carried by the
NANTEN telescope \citep{fukui08}.
Two main peaks in the molecular hydrogen distribution can be seen at R.A., dec = 270.4, -23.4 (cloud N) and at R.A., dec = 270.2, -24.1
(cloud S, see fig. \ref{w28}).
\begin{figure}[b!]
\resizebox{6.5cm}{!}{\includegraphics[clip=true]{w28_gev2.jpg}}
\caption{
\footnotesize
AGILE counts map for W28.
The blue circles indicate the location of the supernova remnant W28, the black contours show the CO intensity emission.
}
\label{w28}
\end{figure}
\begin{figure}[b!]
\resizebox{8cm}{!}{\includegraphics[clip=true]{w28_spec.jpg}}
\caption{
\footnotesize
Combined AGILE and HESS gamma-ray photon spectra for cloud N
(black) and cloud S (red).
The curves represent the gamma-ray spectra estimated (accordling to the model presented in the text) for the two clouds.
}
\label{w28_spec}
\end{figure}
\begin{figure*}[t!]
\resizebox{15cm}{!}{\includegraphics[clip=true]{mevgevtev2.jpg}}
\caption{
\footnotesize
\textit{Upper row}: Predicted gamma-ray emission for energies greater than 400 MeV (left column), 3 GeV (center column) and 400 GeV
(right column) for SNR W28.
\textit{Lower row}: Maps produced from the AGILE, Fermi and HESS observations of SNR W28.
}
\label{mevgevtev}
\end{figure*}
Molecular cloud distribution correlates nicely with the gamma ray observations in both the TeV energy band \citep{aharonian08} and
in the E $>$ 400 MeV energy band observed by AGILE.
However the ratio between the TeV and the multi-MeV emission is significantly different for the cloud N and the cloud S.
In figure \ref{w28_spec} the gamma-ray spectra for the two clouds are shown.
\\The interpretative scenario proposed in \citet{giuliani10} assumes that the N cloud is closer to the CR acceleration site than the S
cloud, considering the energy dependence of the diffusion coefficient.
If protons diffuse in the interstellar medium with a diffusion coefficient given by $D(E) = D_0 \; E^{0.5}$ the resulting proton energy
spectrum is suppressed below a threshold energy $E_t \sim R^4 t^{-2}$ where R is the distance from the acceleration site and \textit{t}
the age of the SNR.
Figure \ref{w28_spec} shows the gamma-ray spectra produced by protons (through neutral pion decay) interacting with the cloud N and S
assuming respectively R = 9 and 4 pc.
\\This scenario can explain also the morphology of the gamma-ray emission seen at different energies ranges (see figure \ref{mevgevtev}).
Assuming that CRs are accelerated in a spherical region (indicated by the blue circle), we evaluated their tridimensional distribution,
N(r,E,t), around the SNR, N(r,E,t), as a function of particle energy and SNR age, solving the diffusion equation :
\[
\frac{dN(r,E,t)}{dt} = D(E) \; \nabla^{2} N + \frac{\partial}{\partial E} \left[ b(E) N\right] + Q(E)
\]
where \textit{b(E)} represents the energy losses, and assuming an impulsive injection with a spectrum
$Q(E) \sim E^{-2.2}$.
Black countours in figure \ref{mevgevtev} show the distribution of targets (molecular hydrogen) as derived by the observations of the NANTEN
telescope.
\\The sky maps in the upper row of figure \ref{mevgevtev} refer to the gamma-ray emission for energies greater than 400 MeV, 3 GeV and
400 GeV expected for an age of 40.000 yrs.
We assumed that gamma-ray emission is produced by p-p collision between accelerated protons and the nuclei of the molecular hydrogen.
In the lower row are shown the maps produced by the AGILE, Fermi and HESS W28 observations.
\section{SNR IC 443}
IC 443 is a SNR lying at a distance of about 1.5 kpc in the Galactic anticenter direction.
Radio, optical and X-rays emission show a shell structure clearly visible, in correspondence of the interaction between SNR and ISM that
produces a shock.
A system of molecular clouds is also associated to the SNR, and an evidence of the interaction is given by the observation of an high
value of the ratio CO (J=2-1)/(J =1-0) \citep{seta98}.
\\A TeV source has been detected both by MAGIC \citep{albert07} and VERITAS \citep{acciari09}.
Thanks to the good angular resolution of the TeV telescopes it was possible to locate the source in a small error box coincident with
the direction of the most massive cloud.
\\AGILE observed gamma-ray emission in this region obtaining an error box which is not compatible with the MAGIC and VERITAS error boxes
\citep{tavani10}.
The different position of the source in the TeV and gamma energy ranges implies a difference in the CR spectrum
or in the target distributions.
A possible interpretation can be given assuming, as in the case of W28, a different distance of the emitting clouds which can lead to a different
spectrum of the accelerated protons seen by the near/far clouds \citep{aharonian06, torres08}.
\begin{table}[h!]
\normalsize
\caption{Middle-aged SNRs seen by~AGILE. Luminosity is given for E $>$ 100 MeV.}
\label{abun}
\begin{center}
\begin{tabular}{|c|c|c|c|}
\hline
SNR & Age & Distance & Luminosity \\
& (\textit{years}) & (\textit{kpc}) & (\textit{erg/s}) \\
\hline
W28 &$ > 35.000 $ & $1.9 $ & $3.3\times10^{34} $ \\
\hline
IC 443 &$ 30.000$ & $1.5 $ & $2.4\times 10^{34}$ \\
\hline
W44 &$ 20.000 $ & $2.8 $ & $6.5\times10^{34} $ \\
\hline
W51C &$ > 20.000 $ & $6.0 $ & $7.3\times10^{34} $ \\
\hline
\end{tabular}
\end{center}
\end{table}
%
\newpage
|
\section{Introduction}
The frequency band structure of artificial periodic dielectrics
formally known as photonic crystals is the electromagnetic (EM)
counterpart of the electronic band structure in ordinary atomic
solids.
Recently, a new analogy between electron and photon states in
periodic structures has been proposed by Raghu and Haldane,
\cite{haldane} namely the one-way chiral edge states in
two-dimensional (2D) photonic-crystal slabs which are similar to
the corresponding edge states in the quantum Hall effect.
\cite{one_way} The photonic chiral edge states are a result of
time-reversal (TR) symmetry breaking which comes about with the
inclusion of gyroelectric/ gyromagnetic material components; these
states are robust to disorder and structural imperfections as long
as the corresponding topological invariant (Chern number in this
case) remains constant.
In certain atomic solids, TR symmetry breaking is not prerequisite
for the appearance of topological electron states as it is the
case in the quantum Hall effect. Namely, when spin-orbit
interactions are included in a TR symmetric graphene sheet, a bulk
excitation gap and spin-filtered edge states emerge
\cite{mele_2005} without the presence of an external magnetic
field, a phenomenon which is known in literature as quantum spin
Hall effect. Its generalization to three-dimensional (3D) atomic
solids lead to a new class of solids, namely, topological
insulators. \cite{ti_papers} The latter possess a
spin-orbit-induced energy gap and gapless surface states
exhibiting insulating behavior in bulk and metallic behavior at
their surfaces.
Apart from topological insulators where the spin-orbit band
structure with TR symmetry defines the topological class of the
corresponding electron states, other topological phases have been
proposed such as topological superconductors (band structure with
particle-hole symmetry), \cite{ts_papers} magnetic insulators
(band structure with magnetic translation symmetry),
\cite{mi_papers} and, very recently, topological crystalline
insulators. \cite{fu_prl} In the latter case the band structure
respects TR symmetry as well as a certain point-group symmetry
leading to bulk energy gap and gapless surface states.
In this work, we propose a photonic analog of a topological
crystalline insulator. Our model photonic system is a 3D crystal
of weakly interacting resonators respecting TR symmetry and the
point-symmetry group associated with a given crystal surface. As a
result, the system possesses an omnidirectional band gap within
which gapless surface states of the EM field are supported. It is
shown that the corresponding photonic band structure is equivalent
to the energy band structure of an atomic topological crystalline
insulator and, as such, the corresponding states are topological
states of the EM field classified by a $Z_{2}$ topological
invariant.
The frequency band structure of photonic crystals whose
(periodically repeated) constituent scattering elements interact
weakly with each other can be calculated by a means which is
similar to the tight-binding method employed for atomic insulators
and semiconductors. Photonic bands amenable to a
tight-binding-like description are e.g., the bands stemming from
the whispering-gallery modes of a lattice of high-index scatterers
\cite{lido} the defect bands of a sublattice of point defects,
within a photonic crystal with an absolute band gap,
\cite{bayindir} the plasmonic bands of a lattice of metallic
spheres \cite{quinten} or of a lattice of dielectric cavities
within a metallic host. \cite{stefanou_ssc}
In the latter case, the frequency band structure stems from the
weak interaction of the surface plasmons of each individual cavity
\cite{stefanou_ssc} wherein light propagates within the crystal
volume by a hopping mechanism. Such type of lattice constitutes
the photonic analog of a topological crystalline insulator
presented in this work whose frequency band structure will be
revealed based on a photonic tight-binding treatment within the
framework of the coupled-dipole method. \cite{cde} The latter is
an exact means of solving Maxwell's equations in the presence of
nonmagnetic scatterers.
\section{Tight-binding description of dielectric cavities in a
plasmonic host}
We consider a lattice of dielectric cavities within a lossless
metallic host. The $i$-th cavity is represented by a dipole of
moment ${\bf P}_{i}=(P_{i;x},P_{i;y},P_{i;z})$ which stems from an
incident electric field ${\bf E}^{inc}$ and the field which is
scattered by all the other cavities of the lattice. This way the
dipole moments of all the cavities are coupled to each other and
to the external field leading to the coupled-dipole equation
\begin{equation}
{\bf P}_{i}= \boldsymbol\alpha_{i}(\omega) [{\bf E}^{inc} +
\sum_{i' \neq i} {\bf G}_{i i'}(\omega) {\bf P}_{i'}].
\label{eq:cde}
\end{equation}
${\bf G}_{i i'}(\omega)$ is the electric part of the free-space
Green's tensor and ${\bf \boldsymbol\alpha}_{i}(\omega)$ is the $3
\times 3$ polarizability tensor of the $i$-th cavity.
Eq.~(\ref{eq:cde}) is a $3N \times 3N$ linear system of equations
where $N$ is the number of cavities of the system. We assume that
the cavities exhibit a uniaxial EM response, i.e., the
corresponding polarizability tensor is diagonal with
$\alpha_{x}=\alpha_{y}=\alpha_{\parallel}$ and
$\alpha_{z}=\alpha_{\perp}$. For strong anisotropy, the cavity
resonances within the $xy$-plane and along the $z$-axis can be
spectrally distinct; thus, around the region of e.g., the cavity
resonance $\omega_{\parallel}$ within the $xy$-plane,
$\alpha_{\perp} \ll \alpha_{\parallel}$ (see appendix). In this
case, one can separate the EM response within the $xy$-plane from
that along the $z$-axis and Eq.~(\ref{eq:cde}) becomes a $2N
\times 2N$ system of equations,
\begin{equation}
{\bf P}_{i}= \alpha_{\parallel}(\omega) [\sum_{i' \neq i} {\bf
G}_{i i'}(\omega) {\bf P}_{i'}]. \label{eq:cde_no_field}
\end{equation}
where we have set ${\bf E}^{inc}={\bf 0}$ since we are seeking the
eigenmodes of the system of cavities. Also, now, ${\bf
P}_{i}=(P_{i;x},P_{i;y})$.
For a particle/cavity of electric permittivity
$\epsilon_{\parallel}$ embedded within a material host of
permittivity $\epsilon_{h}$, the polarizability
$\alpha_{\parallel}$ is given by the Clausius-Mossotti formula
\begin{equation}
\alpha_{\parallel}=\frac{3 V}{4 \pi}
\frac{\epsilon_{\parallel}-\epsilon_{h}}{\epsilon_{\parallel}+
2\epsilon_{h}} \label{eq:cm}
\end{equation}
where $V$ is the volume of the particle/ cavity. For a lossless
plasmonic (metallic) host in which case the electric permittivity
can be taken as Drude-type, i.e., $\epsilon_{h}=1-\omega_{p}^{2} /
\omega^{2}$ (where $\omega_{p}$ is the bulk plasma frequency), the
polarizability $\alpha_{\parallel}$ exhibits a pole at
$\omega_{\parallel}=\omega_{p} \sqrt{2/ (\epsilon_{\parallel}
+2)}$ (surface plasmon resonance). By making a Laurent expansion
of $\alpha_{\parallel}$ around $\omega_{\parallel}$ and keeping
the leading term, we may write
\begin{equation}
\alpha_{\parallel}= \frac{F} {\omega - \omega_{\parallel}} \equiv
\frac{1} {\Omega} \label{eq:a_laurent}
\end{equation}
where $F=(\omega_{\parallel}/2) (\epsilon_{\parallel} -
\epsilon_{h})/ (\epsilon_{\parallel}+2)$. For sufficiently high
value of the permittivity of the dielectric cavity, i.e.,
$\epsilon_{\parallel}
> 10$, the electric field of the surface plasmon is much localized at the surface of the
cavity. As a result, in a periodic lattice of cavities, the
interaction of neighboring surface plasmons is very weak leading
to much narrow frequency bands. By treating such a lattice in a
tight binding-like framework, we may assume that the Green's
tensor ${\bf G}_{i i'}(\omega)$ does not vary much with frequency
and therefore, ${\bf G}_{i i'}(\omega) \simeq {\bf G}_{i
i'}(\omega_{\parallel})$. In this case,
Eq.~(\ref{eq:cde_no_field}) becomes an eigenvalue problem
\begin{equation}
\sum_{i' \neq i} {\bf G}_{i i'}(\omega_{\parallel}) {\bf P}_{i'}=
\Omega {\bf P}_{i} \label{eq:cde_eigen}
\end{equation}
where
\begin{eqnarray}
{\bf G}_{i i'}(\omega_{\parallel})=q_{\parallel}^{3} \Bigl[
C(q_{\parallel} | r_{ii'}|) {\bf
I}_{2} + J(q_{\parallel} | r_{ii'}|) \left(%
\begin{array}{cc}
\frac{x_{ii'}^2}{r_{ii'}^{2}} & \frac{x_{ii'}y_{ii'}}{r_{ii'}^{2}} \\
\frac{x_{ii'}y_{ii'}}{r_{ii'}^{2}} & \frac{y_{ii'}^2}{r_{ii'}^{2}} \\
\end{array}%
\right) \Bigr]. \nonumber \\ \label{eq:g_tensor}
\end{eqnarray}
with ${\bf r}_{ii'}={\bf r}_{i}-{\bf r}_{i'}$,
$q_{\parallel}=\sqrt{\epsilon_{h}}\omega_{\parallel}/c$ and ${\bf
I}_{2}$ is the $2 \times 2$ unit matrix. The form of functions
$C(q_{\parallel} | r_{ii'}|)$, $J(q_{\parallel} | r_{ii'}|)$
generally depends on the type of medium hosting the cavities
(isotropic, gyrotropic, bi-anisotropic, etc).
\cite{eroglu,dmitriev} The Green's tensor of
Eq.~(\ref{eq:g_tensor}) describes the electric interactions
between two point dipoles ${\bf P}_{i}$ and ${\bf P}_{i'}$ each of
which corresponds to a single cavity. The first term of ${\bf
G}_{i i'}$ describes an interaction which does not depend on the
orientation of the two dipoles whilst the second one is
orientation dependent.
\small
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{Fig1.eps}} \caption{(Color
online) (a) Tetragonal crystal with two cavities within the unit
cell. (b) The bulk Brillouin zone and (c) the surface Brillouin
zone corresponding to the (001) surface.} \label{fig1}
\end{figure} \normalsize
For an infinitely periodic system, i.e., a crystal of cavities, we
assume the Bloch ansatz for the polarization field, i.e.,
\begin{equation}
{\bf P}_{i}={\bf P}_{n \beta}=\exp (i {\bf k} \cdot {\bf R}_{n})
{\bf P}_{0 \beta} \label{eq:bloch}
\end{equation}
The cavity index $i$ becomes composite, $i \equiv n \beta$, where
$n$ enumerates the unit cell and $\beta$ the positions of
inequivalent cavities in the unit cell. Also, ${\bf R}_{n}$
denotes the lattice vectors and ${\bf k}=(k_{x},k_{y},k_{z})$ is
the Bloch wavevector. By substituting Eq.~(\ref{eq:bloch}) into
Eq.~(\ref{eq:cde_eigen}) we finally obtain
\begin{equation}
\sum_{\beta'} \tilde{{\bf G}}_{\beta
\beta'}(\omega_{\parallel},{\bf k}) {\bf P}_{0 \beta'}= \Omega
{\bf P}_{0 \beta} \label{eq:cde_eigen_periodic}
\end{equation}
where
\begin{equation}
\tilde{{\bf G}}_{\beta \beta'}(\omega_{\parallel}, {\bf k}) =
\sum_{n'} \exp [i {\bf k} \cdot ({\bf R}_{n}-{\bf R}_{n'})] {\bf
G}_{n \beta; n' \beta'}(\omega_{\parallel}).
\label{eq:green_fourier}
\end{equation}
Solution of Eq.~(\ref{eq:cde_eigen_periodic}) provides the
frequency band structure of a periodic system of cavities.
\small
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{Fig2.eps}}
\caption{Frequency band structure for tetragonal lattice of
resonant cavities within a plasmonic host (see Fig.~\ref{fig1})
corresponding to the Green's tensor of Eq.~(\ref{eq:G_elem}) with
$s^{A}_{1}=-s^{B}_{2}=1.2, s^{A}_{2}=-s^{B}_{2}=0.5,
s'_{1}=2.5,s'_{2}=0.5,s_{z}=2$.} \label{fig2}
\end{figure}
\normalsize
\section{Topological frequency bands}
Since Eq.~(\ref{eq:cde_eigen_periodic}) is equivalent to a
Hamiltonian eigenvalue problem, we adopt the crystal structure of
Ref.~\onlinecite{fu_prl}. Namely, a tetragonal lattice with a unit
cell consisting of two same cavities at inequivalent positions $A$
and $B$ [see Fig.~\ref{fig1}(a)] along the $c$-axis. In this case,
the index $\beta$ in Eq.~(\ref{eq:cde_eigen_periodic}) assumes the
values $\beta=A,B$ for each sublattice (layer) of the crystal. The
above lattice is characterized by the $C_{4}$ point-symmetry
group. In order to preserve the $C_{4}$ symmetry \cite{fu_prl} in
the Green's tensor matrix of Eq.~(\ref{eq:cde_eigen_periodic}) we
assume that the interaction between two cavities within the same
layer (either $A$ or $B$) depends on the relative orientation of
the point dipole in each cavity whilst the interaction between
cavities belonging to adjacent layers is orientation independent.
Also, we take into account interactions up to second neighbors in
both inter- and intra-layer interactions. Taking the above into
account, the lattice Green's tensor assumes the form
\begin{eqnarray}
\tilde{{\bf G}}({\bf k})
=\left(%
\begin{array}{cc}
\tilde{{\bf G}}^{AA}({\bf k}) & \tilde{{\bf G}}^{AB}({\bf k}) \\
\tilde{{\bf G}}^{AB \dagger}({\bf k}) & \tilde{{\bf G}}^{BB}({\bf k}) \\
\end{array}%
\right) \nonumber \\
\end{eqnarray}
where
\begin{eqnarray}
\tilde{{\bf G}}^{\beta \beta}({\bf k})= 2 s^{\beta}_{1}
\left(
\begin{array}{cc}
\cos (k_{x} \alpha) & 0 \\
0 & \cos (k_{y} \alpha) \\
\end{array}
\right)+
\nonumber && \\
2 s^{\beta}_{2}\left(%
\begin{array}{cc}
\cos (k_{x} \alpha)\cos (k_{y} \alpha) & -\sin (k_{x} \alpha)\sin (k_{y} \alpha) \\
-\sin (k_{x} \alpha)\sin (k_{y} \alpha) & \cos (k_{x} \alpha)\cos (k_{y} \alpha) \\
\end{array}%
\right),
\nonumber && \\
\tilde{{\bf G}}^{A B}({\bf k})=[s'_{1}+2 s'_{2} (\cos(k_{x}
\alpha) + \cos (k_{y} \alpha)) +s'_{z} \exp( i k_{z} \alpha)] {\bf
I}_{2}. \nonumber && \\ \label{eq:G_elem}
\end{eqnarray}
\small
\begin{figure}[h]
\centerline{\includegraphics[width=8cm]{Fig3.eps}}
\caption{Frequency band structure for a finite slab $ABAB \cdots
ABB$ of the crystal of Fig.~\ref{fig1} made from 80 bilayers. }
\label{fig3}
\end{figure}
\normalsize
The lattice Green's tensor of Eq.~(\ref{eq:G_elem}) is completely
equivalent to the lattice Hamiltonian of Ref.~\onlinecite{fu_prl}.
$s^{\beta}_{1},s^{\beta}_{2},s'_{1},s'_{2},s_{z}$ in
Eq.~(\ref{eq:G_elem}) generally depend on $q_{\parallel}$, the
lattice constant $a$ and the interlayer distance $c$ but hereafter
will be used as independent parameters. Namely we choose
$s^{A}_{1}=-s^{B}_{2}=1.2, s^{A}_{2}=-s^{B}_{2}=0.5,
s'_{1}=2.5,s'_{2}=0.5,s_{z}=2$. In Fig.~\ref{fig2} we show the
(normalized) frequency band structure corresponding to
Eq.~(\ref{eq:G_elem}) along the symmetry lines of the Brillouin
zone shown in Fig.~\ref{fig1}(b). It is evident that an
omnidirectional frequency band gap exists around $\Omega=0$ which
is prerequisite for the emergence of surface states. In order to
inquire the occurrence of surface states we find the eigenvalues
of the Green's tensor of Eq.~(\ref{eq:G_elem}) in a form
appropriate for a slab geometry. The emergence of surface states
depends critically on the surface termination of the finite slab,
i.e., for different slab terminations different surface-state
dispersions occur (if occur at all). Namely, we assume a finite
slab parallel to the (001) surface (characterized by the $C_{4}$
symmetry group) consisting of 80 alternating $AB$ layers except
the last bilayer which is $BB$, i.e., the layer sequence is $ABAB
\cdots ABB$. The corresponding frequency band structure along the
symmetry lines of the surface Brillouin zone of the (001) surface
[see Fig.~\ref{fig1}(c)] is shown in Fig.~\ref{fig3}. It is
evident that there exist gapless surface states within the band
gap exhibiting a quadratic degeneracy at the $\overline{M}$-point.
In this case, the corresponding doublet of surface states can be
described by an effective theory \cite{chong_prb} similarly to the
doublet states at a point of linear degeneracy (Dirac point).
\cite{sepkhanov}
We note that the equivalence of the Green's tensor ${\bf G}$ with
the atomic Hamiltonian of Ref.~\onlinecite{fu_prl} as well as the
form of the time-reversal $T$ and (geometric) $C_{4}$-rotation $U$
operators for the EM problem \cite{haldane} which are the same as
for spinless electrons, allows to describe the photonic band
structure with the $Z_{2}$ topological invariant $\nu_{0}$
\begin{equation}
(-1)^{\nu_{0}}=(-1)^{\nu_{\Gamma M}} (-1)^{\nu_{A Z}}
\label{eq:z2_def}
\end{equation}
where for real eigenvectors of $\tilde{{\bf G}}({\bf k})$ we have
\cite{fu_prl}
\begin{equation}
(-1)^{{\bf k}_{1} {\bf k}_{2}} = {\rm Pf}[w({\bf k}_{2})]/ {\rm
Pf}[w({\bf k}_{1})] \label{eq:pf_frac}
\end{equation}
and
\begin{equation}
w_{mn}({\bf k}_{i}) = \langle u_{m} ({\bf k}_{i}) | U | u_{n}({\bf
k}_{i}) \rangle. \label{eq:w_def}
\end{equation}
${\rm Pf}$ stands for the Pfaffian of a skew-symmetric matrix,
i.e., ${\rm Pf}[w]^{2}=\det(w)$. Due to the double degeneracy of
the band structure at the four special momenta points $\Gamma, M,
A, Z$ the frequency bands come in doublets. Since frequency
eigenvectors with different eigenfrequencies are orthogonal, all
the inter-pair elements of the $w$-matrix are zero and the latter
is written as:
\begin{equation}
w({\bf k}_{i})=\left(%
\begin{array}{cccc}
w^{1}({\bf k}_{i}) & 0 & 0 & 0 \\
0 & w^{2}({\bf k}_{i}) & 0 & 0 \\
0 & 0 & \ddots & 0 \\
0 & 0 & 0 & w^{N}({\bf k}_{i}) \\
\end{array}%
\right) \label{eq:w_reduced_form}
\end{equation}
where $w^{j}({\bf k}_{i})$ are anti-symmetric $SU(2)$ matrices,
\cite{wang_njp} i.e., $w^{j}({\bf k}_{i})=A_1$ or $A_2$, where
\begin{equation}
A_1= \left(%
\begin{array}{cc}
0 & 1 \\
-1 & 0 \\
\end{array}%
\right), A_2= \left(%
\begin{array}{cc}
0 & -1 \\
1 & 0 \\
\end{array}%
\right). \label{eq:alpha_matr_def}
\end{equation}
In this case, ${\rm Pf}[w({\bf k}_{i})]=w^{1}_{12} w^{2}_{12}
\cdots w^{N}_{12}=\pm 1$. Therefore, $(-1)^{{\bf k}_{1} {\bf
k}_{2}} = \pm 1$ and $\nu_{0}=1$ which ensures the presence of
gapless surface states.
We note that the above analysis relies on the assumption of real
frequency bands. The presence of losses in the constituent
materials renders the frequency bands complex, i.e., the Bloch
wavevector possesses both a real and an imaginary part. However,
even in this case, one can still speak of real frequency bands if
the imaginary part of the Bloch wavevector is at least {\it
hundred} times smaller than the corresponding real part. This a
common criterion used in calculations of the complex frequency
band structure by on-shell electromagnetic solvers such as the
layer-multiple scattering method \cite{comphy} or the
transfer-matrix method. \cite{tmm}
\small
\begin{figure}[h]
\centerline{\includegraphics[width=6cm]{Fig4.eps}} \caption{(Color
online) A possible realization of a photonic structure with
gapless surface states: dielectric particles of square cross
section, joined together with cylindrical coupling elements and
embedded within a 3D network of metallic wires (artificial
plasma).} \label{fig4}
\end{figure}
\normalsize
\section{Blueprint for a photonic topological insulator}
A possible realization of the photonic analogue of topological
insulator in the laboratory is depicted in Fig.~\ref{fig4}. Since
our model system requires dielectric cavities within a homogeneous
plasma, a lattice of nano-cavities formed within a homogeneous
Drude-type metal, e.g., a noble metal (Au, Ag, Cu), would be the
obvious answer. \cite{stefanou_ssc} However, the plasmon bands are
extremely lossy due to the intrinsic absorption of noble metals in
the visible regime. A solution to this would be the use of an {\it
artificial} plasmonic medium operating in the microwave regime
where metals are perfect conductors and losses are minimal.
Artificial plasma can be created by a 3D network of thin metallic
wires of a few tens of $\mu$m in diameter and spaced by a few mm.
\cite{art_plasma} A lattice of dielectric particles within an
artificial plasma can be modelled with the presented tight-binding
Green's tensor. Since the interaction among first and second
neighbors within the same bilayer ($A$ or $B$) should depend on
the dipole orientation (in order to preserve the $C_{4}$
symmetry), the dielectric particles in each layer are connected
with cylindrical waveguiding elements (different in each layer $A$
or $B$ - see Fig.~\ref{fig4}). In contrast, between two successive
bilayers there are no such elements since interactions between
dipoles belonging to different layers should be independent of the
dipole orientations (s orbital-like). Another advantage of
realizing the photonic analog in the microwave regime is the
absence of nonlinearities in the EM response of the constituent
materials since photon-photon interactions may destroy the
quadratic degeneracy of the surface bands in analogy with
fermionic systems. \cite{sun}
Finally, we must stress that a photonic topological crystalline
insulator can be also realized with purely dielectric materials if
the host medium surrounding the cavities is not a plasmonic medium
but a photonic crystal with an absolute band gap: the cavities
would be point defects within the otherwise periodic photonic
crystal and the tight-binding description would be still
appropriate. \cite{bayindir,k_fang} In this case, Maxwell's
equations lack of any kind of characteristic length and the
proposed analog would be realized in any length scale.
\section{Conclusions}
In conclusion, a 3D lattice of weakly interacting cavities
respecting TR symmetry and a certain point-group symmetry
constitutes a photonic analog of a topological crystalline
insulator by demonstrating a spectrum of gapless surface states. A
possible experimental realization would be a 3D lattice of
dielectric particles within a continuous network of thin metallic
wires with a plasma frequency in the GHz regime.
This work has been supported by the European Community's Seventh
Framework Programme (FP7/2007-2013) under Grant Agreement No.
228455-NANOGOLD (Self-organized nanomaterials for tailored optical
and electrical properties).
|
\section{\label{sec:1}Introduction}
\setcounter{equation}{0}
The idea of the multidimensionality of our Universe has been attracting continuous interest for many years. It takes its origin from the pioneering papers by Th. Kaluza
and O. Klein \cite{KK}, and now the most self-consistent modern theories of unification such as superstrings, supergravity and M-theory are constructed in spacetime with
extra dimensions \cite{Polchinski}. In Kaluza-Klein models, our spacetime is effectively four-dimensional due to compactness and smallness of the extra dimensions
(internal spaces). The size of the extra dimensions is restricted by the electroweak scales $10^{-17}$ cm. However, our spacetime can be effectively four-dimensional
even in the case of infinite extra dimensions. This interesting scenario is realized in recently proposed brane world models (see, e.g., the reviews \cite{Rub,Barv}).
Here, matter fields from the Standard Model are trapped to a three-dimensional submanifold (brane) embedded in the fundamental multidimensional space (bulk), but
gravity may move in the bulk. Localization of massless gravitons on a brane results in effective four-dimensional Einstein gravity in the low energy limit. Certainly,
large and infinite extra dimensions are potentially detectable. This was one of the main reasons for the great interest in this scenario. Therefore, it is very important
to suggest experiments which can reveal such extra dimensions.
In our paper, we consider the scenario that was first proposed in \cite{RS2}. Here, the brane is embedded in the five-dimensional anti-DeSitter spacetime, which allows
the extra dimension to be infinite. A negative bulk cosmological constant $\Lambda_5$ and a brane tension $\sigma$ are fine tuned to each other. Clearly, this is a very
simplified scenario. However, it gives a possibility to reveal some general features of the brane world models, in particular, the localization of the massless graviton
on the brane that restores the Newtonian limit on the brane at large distances from the gravitating matter source. It was shown \cite{RS2} that at distances greater than
a curvature scale of anti-DeSitter spacetime $r\gg l\sim |\Lambda_5|^{-1/2}$, the gravitational potential takes an approximate form with a cubic additive $\sim 1/r^3$ to
the usual Newtonian potential $\sim 1/r$. This approximate solution is much simpler than the exact one that makes the investigation of the effects of the extra dimension
much easier. In some papers (see, e.g., \cite{approx1,approx2}) this approximation was used to calculate the gravitational interaction between gravitating test bodies of
different geometrical form.
But we should analyze the difference between the approximate and exact solutions to find out where the application of the approximate solution is appropriate. This is
one of the main motivations of this work. To perform such analysis, we obtain two types of solutions (approximate and exact) for gravitational potentials and
accelerations of test bodies in these potentials for different geometrical configurations. Then, we apply these formulas to the most interesting for experiments case of
gravitational interaction between two massive spheres. We calculate approximate and exact corrections to the Newton's gravitational force and show that the difference
between relative force corrections for the approximate and exact cases increases with the parameter $l$ (for the fixed distance $r$ between centers of the spheres). On
the other hand, this difference increases with decreasing of the distance between the centers of the spheres (for the fixed curvature scale parameter $l$). The relative
force corrections also allow us to get the experimental constraint on the curvature scale parameter: $l\lesssim 10\, \mu$m. To get it, we use the results of the
table-top inverse square law experiments for the measurements of the Newton's gravitational constant. This is one of the main results of our paper.
The paper is structured as follows. In section 2 we describe briefly the Randall-Sundrum model with one brane. Here, we consider non-relativistic limit of this model and
present approximate and exact solutions for the gravitational potential on the brane. These formulas are applied to some practical problems in section 3 to get
approximate and exact expressions for the gravitational potential and acceleration of a point mass for these problems. In section 4 we investigate the gravitational
interaction of two spherical shells. Then, in section 5 we compare the relative corrections to the gravitational force between two spheres in approximate and exact
cases. Here, we also get the constraint on the curvature scale parameter in the Randall-Sundrum model. A brief discussions of the obtained results is presented in the
concluding section 6.
\section{\label{sec:2}Non-relativistic limit of Randall-Sundrum model}
\setcounter{equation}{0}
The one-brane Randall-Sundrum metrics is \cite{RS2}
\be{2.1}
ds^2=\exp\left(-\frac{2|\xi|}{l}\right)\eta_{\mu\nu}dx^{\mu}dx^{\nu}-d\xi^2\ ,
\ee
where $\eta_{\mu\nu}$ is the flat four-dimensional spacetime metrics and the parameter $l$ is defined via the 5-dimensional cosmological constant:
\be{2.2}
\Lambda_5=-\frac6{l^2}\, ,
\ee
i.e. $l$ is the curvature scale of 5-dimensional anti-DeSitter spacetime. The brane is embedded in this spacetime at $\xi=0$ and has fine tuned tension
\be{2.3} \sigma=\frac{3c^4}{4\pi G_5 l}\, , \ee
where $G_5$ is the 5-dimensional gravitational constant. In one-brane Randall-Sundrum scenario, the extra dimensions is infinite: $\xi \in (-\infty,+\infty)$.
Now, we want to probe this model with the help of the gravitational terrestrial experiments, e.g., the inverse square law experiments. Certainly, this is the case of
non-relativistic limit of the model. In this limit, we need to get the gravitational potential $\varphi (r)$ on the brane.
Following, e.g., the calculations in \cite{Barv}, we obtain
\be{2.4}
\varphi(r)=-\frac{G_5 m}{r l}-\frac{G_5 m}{r}\int\limits_0^{\infty}d\tilde m \varphi^2_{\tilde m}(l)\exp(-\tilde m r)\, ,
\ee
where $r$ is the magnitude of the radius vector on the brane and
\be{2.5}\varphi_{\tilde m}(l)=\left(\frac{\tilde m l}{2}\right)^{1/2}\frac{Y_1(\tilde m l)J_2(\tilde m l)-J_1(\tilde m l)Y_2(\tilde m l)}{\left(J^2_1(\tilde m
l)+Y^2_1(\tilde m l)\right)^{1/2}}\, , \ee
where $J$ and $Y$ are Bessel functions of the first and second kinds, respectively. In the short and long distance limits the equation \rf{2.4} reads, respectively:
\be{2.6} \varphi(r)\approx-\frac{G_5 m}{rl}-\frac{G_5 m}{r}\frac{1}{\pi r}
\approx-\frac{G_5 m }{\pi r^2}\, ,\ \ \ \ r\ll l\,
\ee
and
\be{2.7}
\varphi(r)\approx-\frac{G_N m}{r}\left(1+\frac{\alpha}{r^2}\right)\ , \ \ \ r\gg l\, ,
\ee
where we have introduced the Newton's gravitational constant
\be{2.8}
G_N=\frac{G_5}{l}
\ee
and the parameter\footnote{\label{alpha}It is worth noting that in the pioneering paper \cite{RS2} $\alpha = l^2$. The brane-bending effect \cite{GT} gives
$\alpha=2l^2/3$. In the paper \cite{JKP}, the authors have pointed out that different schemes of regularization result in different values of $\alpha$. In our paper we
follow calculations in \cite{Barv}, where $\alpha=l^2/2$.}
\be{2.9}
\alpha = \frac{l^2}{2}\, .
\ee
Obviously, \rf{2.6} corresponds to the strong deviation from the Newtonian gravity but the formula \rf{2.7} describes the smooth transition to the Newtonian limit. The
exact expression \rf{2.4} (the solid line) and its asymptotes \rf{2.6} and \rf{2.7} (the short-dashed and long-dashed lines, respectively) are depicted on figure 1.
Here, we introduce the dimensionless distance argument $\eta=r/l$ and dimensionless potentials $\tilde \varphi (\eta) =\varphi (r)/(G_Nm/l)$.
\begin{figure}
\center
\includegraphics[width=3.0in,height=2.2in]{phiotr.eps}\\
\caption {The gravitational field potential \rf{2.4} (the solid line) and its asymptotes \rf{2.6} and \rf{2.7} (the short-dashed and long-dashed lines, respectively).}
\end{figure}
\section{\label{sec:3}Applications}
\setcounter{equation}{0}
Now, we want to apply the obtained formulas to terrestrial gravitational experiments. Obviously, the gravitational field on the Earth should not considerably differ from
the Newtonian one. Therefore, we should use either the exact expression \rf{2.4} or the approximate formula \rf{2.7}. Therefore, we shall get two classes of solutions:
exact and approximate, respectively. Obviously, the approximate formula \rf{2.7} looks much more simple. However, it is necessary to check the deviation of expressions
based on it from the exact ones for real gravitational experiments. This is one of the main aims of the paper.
\subsection{Infinitesimally thin shell}
Let us consider first an infinitesimally thin shell of the mass $m=4\pi R^2\sigma$, where $R$ and $\sigma$ are the radius and the surface mass density of the shell.
Then, the gravitational potential of this shell in a point with the radius vector $\bf r$ (from the center of the shell) for the approximate solution is
\be{3.1} \varphi(r>R)=-\frac{G_N m}{r}\left[1+\frac{\alpha}{r^2-R^2}\right]\ee
and
\be{3.2}
\varphi(r<R)=-\frac{G_N m}{R}\left[1+\frac{\alpha}{R^2-r^2}\right]\, .
\ee
Obviously, these expressions are divergent when $r\rightarrow R$: $\varphi(r\gtrless R)\rightarrow-G_N m\alpha/\left[2R^2|r-R|\right] \rightarrow -\infty$. In the case
of the exact solution we get
\ba{3.3}
\fl \varphi (r>R)=-\frac{G_N m}{r}\left[1+\frac{l^2}{2R}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde
ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\sinh(\tilde m R)e^{-\tilde m r}\right]\nn \\
\phantom{}
\ea
and
\ba{3.4}
\fl \varphi(r<R)=-\frac{G_N m}{R}\left[1+\frac{l^2}{2r}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde
ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\sinh(\tilde m r) e^{-\tilde m R} \right].\nn \\
\phantom{}
\ea
It is not difficult to verify that these exact solutions are also divergent when $r\to R$.
Formulas \rf{3.2} and \rf{3.4} demonstrate that inside of the shell the gravitational potential is not a constant. Thus, a test body undergoes an acceleration (see
\rf{3.6} and \rf{3.8} below) in contract to the Newtonian case, i.e. the Birkhoff's theorem is violated. The acceleration outside and inside of the shell is
\be{3.5}
-\frac{d\varphi}{dr}(r>R)=-\frac{G_N m}{r^2}\left[1+\alpha\frac{3r^2 - R^2}{(r^2-R^2)^2}\right]
\ee
and
\be{3.6}
-\frac{d\varphi}{dr}(r<R)=\frac{G_N m}{R}\frac{2\alpha r}{(R^2-r^2)^2}\, ,
\ee
which is divergent when $r\rightarrow R$: $-\frac{d\varphi}{dr}(r\gtrless R)\rightarrow \mp \frac{G_N m\alpha}{2R^2(r-R)^2} \rightarrow \mp \infty$ (the upper and lower
signs correspond to $r>R$ and $r<R$, respectively).
In the case of exact solutions we have
\ba{3.7}
\fl &-&\frac{d\varphi}{dr} (r>R)=-\frac{G_N m}{r^2}\\
\fl &-&\frac{G_N m l^2}{2Rr^2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde
ml)Y_2(\tilde ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\sinh(\tilde m R)(1+\tilde m r)e^{-\tilde m r} \nonumber
\ea
and
\ba{3.8}
\fl &-&\frac{d\varphi}{dr}(r<R)\\
\fl &=&\frac{G_N ml^2}{2Rr^2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde
ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2(\tilde m r\cosh(\tilde m r)-\sinh(\tilde m r))e^{-\tilde m R}. \nonumber \ea
The exact solutions \rf{3.7} and \rf{3.8} are also divergent for $r\rightarrow R$.
\subsection{Spherical shell of finite thickness}
Here, we consider a spherical shell of the inner radius $R_1$ and the outer radius $R_2$ and the mass $m=(4\pi \rho/3)\left(R_2^3-R_1^3\right)$ with a constant volume
density $\rho$\footnote{\label{limit}It is clear that the limit $R_2 \rightarrow R_1$ is incorrect because, for constant $\rho$, it results in vanishing $m$ and, vice
versa, for fixed $m$ the volume density $\rho$ goes to infinity. Therefore, such a naive limit does not provide us the correct transition to the formulas from the
previous subsection.}. For this geometry, the approximate gravitational potential reads
\be{3.9}
\fl \varphi(r>R_2)=-\frac{G_N m}{r}\left[1+\frac{3\alpha}{R_2^3-R_1^3}\left(R_1-R_2+\frac{r}{2}\ln\frac{(r+R_2)(r-R_1)}{(r-R_2)(r+R_1)}\right)\right]
\ee
and
\be{3.10}
\fl \varphi(r<R_1)=-2\pi G_N\rho\left(R_2^2-R_1^2+\alpha\ln\frac{R_2^2-r^2}{R_1^2-r^2}\right)\, .
\ee
These expressions are logarithmically divergent in the vicinity of $R_1$ and $R_2$: $\varphi(r>R_2)\rightarrow \frac{3G_N m\alpha}{2\left(R_2^3-R_1^3\right)}\ln(r-R_2)
\rightarrow -\infty$ for $r\rightarrow R_2$ and $\varphi(r<R_1)\rightarrow 2\pi G_N\rho\alpha\ln(R_1-r) \rightarrow -\infty $ for $r\rightarrow R_1$.
For the exact solution we get
\ba{3.11}
\fl &{}& \varphi(r>R_2)=-\frac{G_N m}{r}\left\{1+\frac{l^2}{2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde
ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\frac1{\tilde m^2}e^{-\tilde m r}\right.\nonumber\\
\fl &{}&\left.\times\frac3{(R_2^3-R_1^3)}\left[R\tilde m\cosh(\tilde m R)-\sinh(\tilde mR)\right]|_{R_1}^{R_2}\right\}
\ea
and
\ba{3.12}
\fl &{}&\varphi(r<R_1)=-4\pi G_N\rho\left\{\frac{R^2}{2}-\frac{l^2}{2r}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde
ml)Y_2(\tilde ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\frac1{\tilde m^2}\sinh(\tilde m r)\right.\nonumber\\
\fl &{}&\times\left.(\tilde m R +1)e^{-\tilde m R}\phantom{\int}\right\}|_{R_1}^{R_2}\, .
\ea
In contrast to the approximate formulas \rf{3.9} and \rf{3.10}, these exact expressions are convergent in the limits $r \rightarrow R_1,R_2$.
The acceleration of a test body outside and inside of the shell is
\be{3.13}
\fl -\frac{d\varphi}{dr}(r>R_2)=-\frac{G_N
m}{r^2}\left[1+\frac{3\alpha}{R_2^3-R_1^3}\left(R_1-R_2+r^2\frac{(R_2-R_1)(r^2+R_1R_2)}{(r^2-R_1^2)(r^2-R_2^2)}\right)\right]\quad
\ee
and
\be{3.14}
\fl -\frac{d\varphi}{dr}(r<R_1)=4\pi G_N\rho\alpha r\frac{R_2^2-R_1^2}{(R_2^2-r^2)(R_1^2-r^2)}\, .
\ee
These formulas are divergent in the limits $r \rightarrow R_1,R_2$: $-\frac{d\varphi}{dr}(r>R_2)\rightarrow -\frac{3G_N m\alpha}{2\left(R_2^3-R_1^3\right)(r-R_2)}
\rightarrow -\infty$ for $r\rightarrow R_2$ and $-\frac{d\varphi}{dr}(r<R_1)\rightarrow \frac{2\pi G_N\rho\alpha}{R_1-r} \rightarrow +\infty$ for $r\rightarrow R_1$.
For exact solutions we obtain
\ba{3.15}
\fl &-&\frac{d\varphi}{dr}(r>R_2)=-\frac{G_N m}{r^2}-\frac{G_N ml^2}{2r^2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde
ml)Y_2(\tilde
ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\nonumber\\
\fl &\times&\frac1{\tilde m^2}e^{-\tilde m r}\left(1+\tilde m r\right)\frac3{(R_2^3-R_1^3)}\left[R\tilde m\cosh(\tilde m R)-\sinh(\tilde mR)\right]|_{R_1}^{R_2}
\ea
and
\ba{3.16}
\fl &-&\frac{d\varphi}{dr}(r<R_1)=-\frac{2\pi G_N\rho l^2}{r^2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde
ml)}{(J_1^2(\tilde
ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\nonumber\\
\fl &\times& \frac1{\tilde m^2}(\tilde m r\cosh(\tilde m r)-\sinh(\tilde m r)) (\tilde m R+1)e^{-\tilde m R}|_{R_1}^{R_2}\, .
\ea
These integrals are divergent in the vicinity of $R_1$ and $R_2$.
\subsection{Sphere}
Obviously, all formulas for a sphere of the radius $R$ and the mass $m=4\pi \rho R^3 /3$ with a constant volume density $\rho$ can be easily obtained from the equations
\rf{3.9}, \rf{3.11}, \rf{3.13} and \rf{3.15} with the help of the evident substitutions: $R_1=0$ and $R_2\equiv R$.
\section{\label{sec:4}Gravitational interaction of two spherical shells}
\setcounter{equation}{0}
Let us consider now two spherical shells with radii $R_2>R_1$ and the mass $m=(4\pi \rho/3)\left(R_2^3-R_1^3\right)$ for the first shell and radii $R'_2>R'_1$ and the
mass $m'=(4\pi \rho'/3)\left(R_2^{'3}-R_1^{'3}\right)$ for the second shell. Then, the potential energy of gravitational interaction between these shells for the
approximate solution reads
\ba{4.1}
\fl U(r)&=&-\frac{G_N mm'}{r}-\frac{2\pi^2G_N\rho\rho'\alpha}{r}
\left\{\left[-\frac1{12}r^4+\frac1{2}r^2\left(R'^2+R^2\right)+\frac1{4}\left(R'^2-R^2\right)^2\right]\right.\nn \\
\fl &\times& \ln\frac{r^2-(R'+R)^2}{r^2-(R'-R)^2}
+\frac{2}{3}r\left[R'^3\ln\frac{(r+R)^2-R'^2}{(r-R)^2-R'^2}+R^3\ln\frac{(r+R')^2-R^2}{(r-R')^2-R^2}\right]\nn \\
\fl &-&\left.\frac1{3}r^2R'R-R'^3R-R'R^3\right\}|_{R=R_1}^{R=R_2}|_{R'=R'_1}^{R'=R'_2}\, ,
\ea
where $r\geqslant R_2+R'_2$ is the distance between the centers of the shells and $f(R,R')|_{R=R_1}^{R=R_2}|_{R'=R'_1}^{R'=R'_2}=f(R_2,R'_2) -f(R_2,R'_1) -f(R_1,R'_2)
+f(R_1,R'_1)$.
In the case of the exact solution we get
\ba{4.2}
\fl U(r)&=&-\frac{G_N m m'}{r}-\frac{G_N m m'}{r}\, \frac9{(R_2^3-R_1^3)(R_2^{'3}-R_1^{'3})} \nonumber\\
\fl &\times& \frac{l^2}{2}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde ml)}{(J_1^2(\tilde ml)+Y_1^2(\tilde
ml))^{1/2}}\right]^2 \frac1{\tilde m^5}e^{-\tilde m r}\nonumber\\
\fl &\times& \left[\tilde m R\cosh(\tilde m R)-\sinh(\tilde m R)\right]|_{R_1}^{R_2}\cdot\left[\tilde m R\cosh(\tilde m R)-\sinh(\tilde m R)\right]|_{R'_1}^{R'_2}\, .
\ea
The additional analysis shows that both of these expressions \rf{4.1} and \rf{4.2} are convergent in the limit $r\rightarrow R_2+R'_2$.
With the help of these formulas, we can obtain the absolute value of the gravitational force between two shells:
\be{4.3}
F(r)=\frac{dU}{dr} = \frac{G_N m m'}{r^2}(1+\delta_F)\, ,
\ee
where $\delta_F$ defines the relative deviation from the Newtonian expression $G_N m m'/r^2$. For the approximate and exact solutions we have, respectively:
\ba{4.4}
\fl \delta_F&=&-\frac{9\alpha}{8\left(R_2^3-R_1^3\right)\left(R_2'^3-R_1'^3\right)}\left\{\ln\frac{r^2-(R'+R)^2}{r^2-(R'-R)^2}
\left[-\frac1{4}r^4+\frac1{2}r^2\left(R'^2+R^2\right)\right.\right.\nn \\
\fl &-&\left.\left.\frac1{4}
\left(R'^2-R^2\right)^2\right]
-r^2R'R+R'^3R+R'R^3\right\}|_{R=R_1}^{R=R_2}|_{R'=R'_1}^{R'=R'_2}
\ea
and
\ba{4.5}
\fl \delta_F&=&\frac{9l^2}{2\left(R_2^3-R_1^3\right)\left(R_2^{'3}-R_1^{'3}\right)} \int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde
ml)-J_1(\tilde ml)Y_2(\tilde ml)}{(J_1^2(\tilde
ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\frac 1{\tilde m^5}e^{-\tilde m r}\nonumber\\
\fl &\times& (1+\tilde m r)\left[\tilde m R\cosh(\tilde m R)-\sinh(\tilde m R)\right]|_{R_1}^{R_2}
\cdot\left[\tilde m R'\cosh(\tilde m R')-\sinh(\tilde m R')\right]|_{R'_1}^{R'_2}\, .\nn \\
\fl
\ea
These relative corrections $\delta_F$ are also convergent in the limit $r\rightarrow R_2+R'_2$. In the limit of large separation between the shells $r\gg
R_{1,2},R'_{1,2}$ we obtain from \rf{4.4} $\delta_F=3\alpha/r^2 + O(1/r^3)$.
\section{\label{sec:5}Constraints}
\setcounter{equation}{0}
The obtained above formulas can be used for the experimental restrictions on the parameters of the model. In our case, it is the curvature scale $l$. To get it, we can
use the inverse square law experiments for two spheres. The potential energy of interaction and the gravitational force between two spheres follow from the previous
section with the help of the substitutions: $R_1=R'_1=0$ and $R_2\equiv R,R'_2=R'$. For example, the relative corrections to the gravitational force in approximate and
exact cases read, respectively:
\ba{5.1}
\fl \delta_F&=&-\frac{9\alpha}{8R^3R'^3}\left\{\ln\frac{r^2-(R'+R)^2}{r^2-(R'-R)^2} \left[-\frac1{4}r^4+\frac1{2}r^2\left(R'^2+R^2\right)-\frac1{4}
\left(R'^2-R^2\right)^2\right]\right.\nonumber\\
\fl&-& \left.r^2R'R+R'^3R+R'R^3\right\}
\ea
and
\ba{5.2}
\fl \delta_F&=&\frac{9l^2}{2R^3R^{'3}}\int\limits_0^{\infty}d\tilde m \left[\frac{Y_1(\tilde ml)J_2(\tilde ml)-J_1(\tilde ml)Y_2(\tilde ml)}{(J_1^2(\tilde
ml)+Y_1^2(\tilde ml))^{1/2}}\right]^2\frac 1{\tilde m^5}e^{-\tilde m r}(1+\tilde m r)\nonumber\\
\fl &\times& \left[\tilde m R\cosh(\tilde m R)-\sinh(\tilde m R)\right]\cdot\left[\tilde m R'\cosh(\tilde m R')-\sinh(\tilde m R')\right]\, .
\ea
We remind that $\alpha=l^2/2$ (see the equation \rf{2.9}). For definiteness, we shall use the parameters of the spheres from the Moscow Cavendish-type experiment
\cite{Moscow}: $R_1\approx 0.087$ cm for a platinum ball with the mass $m_1=59.25\times10^{-3}$ g, $R_2\approx0.206$~cm for a tungsten ball with the mass
$m_2=706\times10^{-3}$ g and the distance between their centers $r=0.3773$ cm.
It is clear that the use of the approximate solution for the gravitational interaction force makes the calculations much easier. But we should analyze the distinction
between the approximate and exact solutions to find out where the application of the approximate solution is appropriate. The difference between the approximate and
exact solutions for the relative force corrections is shown on figures 2 and 3. These figures demonstrate that the difference between these solutions increases with the
parameter $l$ (for the fixed distance $r$ between centers of the spheres) (figure 2). On the other hand, this difference increases with decreasing of the distance
between the centers of the spheres (for the fixed curvature scale parameter $l$) (figure 3).
\begin{figure}
\center
\includegraphics[width=3.0in,height=2.0in]{deltaotl.eps}\\
\caption {Relative gravitational force corrections \rf{5.2} (the solid line) and \rf{5.1} (the dashed line) as functions of the curvature scale parameter $l$ in the case
of the distance between the centers of the spheres $r=0.3773$ cm.}
\end{figure}
\begin{figure}
\center
\includegraphics[width=3.0in,height=2.2in]{deltaotr.eps}\\
\caption {Relative gravitational force corrections \rf{5.2} (the solid line) and \rf{5.1} (the dashed line) as functions of the distance between the centers of the
spheres in the case $l=10^{-1}$ cm.}
\end{figure}
Now, we want to estimate the curvature scale parameter $l$ with the help of our formulas \rf{5.1} and \rf{5.2}. To get it, we can use the value of the Newton's
gravitational constant $G_N$.
As it follows from figure 2 in the 'CODATA Recommended Values of the Fundamental Constants: 2006', the most precise values of $G_N$ were obtained in the University
Washington and the University Z\"urich experiments \cite{13,14}. They are $G_N / 10^{-11} {\mbox{m}}^3{\mbox{kg}}^{-1}{\mbox{s}}^{-2}=6.674\, 215\pm 0.000\, 092$ and
$6.674\, 252\pm 0.000\, 124$, respectively. The relative errors $\triangle G_N/G_N$ show the accuracy of the measurements of the gravitational constant in the inverse
square law experiments. If the correction $\delta_F$ due to the extra dimension is greater than these values, then we can detect the deviation from the Newton's law. Up
to now, there is no experimental evidence for such deviations. Therefore, the relation $|\triangle G_N/G_N | =\delta_F$ gives the upper limit for $\delta_F$. In turn,
the equations \rf{5.1} and \rf{5.2} show that $\delta_F \sim l^2$. Therefore, from these equations we can get the upper limit for $l$, substituting there for
definiteness values for the radii of the spheres and the separation between them from the Moscow experiment. Thus, for the Washington and Z\"urich experiments, in the
case of the approximate formula \rf{5.1} we get respectively 9.067$\, \mu$m and 10.527$\, \mu$m and in the case of the exact formula \rf{5.2} we obtain respectively
9.070$\, \mu$m and 10.531$\, \mu$m.
Of course, we get rather rough estimates for the upper limit of $l$. Anyway, we think that it gives more or less correct value of the order of magnitude of $l$ in the
Randall-Sundrum model with one brane: $l\lesssim 10\, \mu$m. Figure 2 shows that for such values of $l$ the difference between the approximate and exact formulas is
negligible. It is worth noting that close constraints were found in the table-top inverse square law experiments \cite{ISLex} and from astrophysical observations
\cite{BH1}-\cite{BH5}.
\section{Conclusion}
In our paper we have considered the one-brane Randall-Sundrum model. In the weak-field limit, we obtained the approximate and exact expressions for gravitational
potentials and accelerations of test bodies in these potentials for different geometrical configurations. Some of these approximate formulas were already known (see,
e.g., \cite{approx1,approx2}), but the exact ones were found for the first time. We applied these equations for calculation of the gravitational interaction between two
spherical shells of finite thickness that can be easily reduced to the case of spheres. Then, we found the approximate and exact expressions for the relative force
corrections to the Newton's gravitational force between two massive spheres. It is clear that the use of the approximate solution makes the calculations much easier. But
we should analyze the difference between the approximate and exact solutions to find out where the application of the approximate solution is appropriate. We found that
the difference between relative force corrections for the approximate and exact cases increases with the parameter $l$ (for the fixed distance $r$ between centers of the
spheres). On the other hand, this difference increases with decreasing of the distance between the centers of the spheres (for the fixed curvature scale parameter $l$).
Using the results of the table-top Cavendish-type experiments measuring the Newton's gravitational constant, from the equations for the relative force corrections we got
the upper limit for the curvature scale parameter $l\lesssim 10\, \mu$m in the Randall-Sundrum model. For these values of $l$, the difference between the approximate and
exact solutions is negligible.
\ack This work was supported in part by the "Cosmomicrophysics" programme of the Physics and Astronomy Division of the National Academy of Sciences of Ukraine.
\section*{References}
|
Subsets and Splits